You are on page 1of 41

BI85CH10-Qi ARI 9 May 2016 9:29

ANNUAL
REVIEWS Further
Click here to view this article's
online features:
• Download figures as PPT slides
• Navigate linked references
• Download citations
• Explore related articles
CRISPR/Cas9 in Genome
• Search keywords
Editing and Beyond
Haifeng Wang,1 Marie La Russa,1,2 and Lei S. Qi1,3,4
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

1
Department of Bioengineering, Stanford University, Stanford, California 94305;
email: hfgwang@stanford.edu, mlarussa@stanford.edu, stanley.qi@stanford.edu
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

2
Biomedical Sciences Graduate Program, University of California, San Francisco,
California 94158
3
Department of Chemical and Systems Biology, Stanford University, Stanford, California 94305
4
Chemistry, Engineering and Medicine for Human Health (ChEM–H), Stanford University,
Stanford, California 94305

Annu. Rev. Biochem. 2016. 85:227–64 Keywords


First published online as a Review in Advance on
dCas9, Cas9 structure, gene regulation, epigenetic regulation, genomic
April 25, 2016
imaging, CRISPR applications
The Annual Review of Biochemistry is online at
biochem.annualreviews.org Abstract
This article’s doi:
The Cas9 protein (CRISPR-associated protein 9), derived from type II
10.1146/annurev-biochem-060815-014607
CRISPR (clustered regularly interspaced short palindromic repeats) bac-
Copyright  c 2016 by Annual Reviews.
terial immune systems, is emerging as a powerful tool for engineering the
All rights reserved
genome in diverse organisms. As an RNA-guided DNA endonuclease, Cas9
can be easily programmed to target new sites by altering its guide RNA
sequence, and its development as a tool has made sequence-specific gene
editing several magnitudes easier. The nuclease-deactivated form of Cas9
further provides a versatile RNA-guided DNA-targeting platform for regu-
lating and imaging the genome, as well as for rewriting the epigenetic status,
all in a sequence-specific manner. With all of these advances, we have just
begun to explore the possible applications of Cas9 in biomedical research and
therapeutics. In this review, we describe the current models of Cas9 function
and the structural and biochemical studies that support it. We focus on the
applications of Cas9 for genome editing, regulation, and imaging, discuss
other possible applications and some technical considerations, and highlight
the many advantages that CRISPR/Cas9 technology offers.

227
BI85CH10-Qi ARI 9 May 2016 9:29

Contents
INTRODUCTION: TOOLS FOR PROGRAMMABLE GENOME EDITING,
TARGETING, AND REGULATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
CRISPR/CAS9: A GIFT FROM MOTHER NATURE. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
CRISPR: An Adaptive Immune Mechanism in Bacteria and Archaea. . . . . . . . . . . . . . . 231
Repurposing Cas9 for Sequence-Specific Genomic Targeting . . . . . . . . . . . . . . . . . . . . . 232
Diversity of Cas9 Orthologs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
Engineered Variants of the Cas9 Nuclease Domains: Nickase Cas9 (nCas9)
and Nuclease-Deactivated Cas9 (dCas9) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Other RNA-Guided Endonucleases in the CRISPR Systems . . . . . . . . . . . . . . . . . . . . . . 233
STRUCTURE OF Cas9 AND PROPOSED WORKING MODEL FOR GUIDE
RNA BINDING AND TARGET DNA CLEAVAGE . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Bilobed Structure of Cas9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234


DNA Targeting Specificity of Cas9: Protospacer-Adjacent
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Motif (PAM) Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234


DNA Targeting Specificity of Cas9: Guide RNA–Target DNA
Base-Pairing Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
Interactions between the Single Guide RNA–DNA Duplex and Cas9 . . . . . . . . . . . . . . 235
Orthogonality of Cas9–Single Guide RNA Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
Proposed Working Model for Guide RNA Binding and Target DNA Cleavage . . . . 236
Cas9 FOR GENOME EDITING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
Mechanism of Genome Editing: DNA Cleavage Followed by DNA Repair . . . . . . . . 236
Applications of Cas9-Mediated Genome Editing for Studying Gene Function,
Disease Modeling, and Gene Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
Applications of Cas9-Mediated Genome Editing for Genome-Wide
Functional Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
Challenges in Cas9-Mediated Genome Editing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
Responsible Use of Cas9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Cas9 FOR GENE REGULATION: CRISPR INTERFERENCE (CRISPRi)
AND CRISPR ACTIVATION (CRISPRa) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Nuclease-Deactivated Cas9 (dCas9): A Programmable Platform for
Sequence-Specific Gene Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Multimodal CRISPRi/a Function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Strategies to Improve CRISPRa Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Advantages of CRISPRi/a and Their Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
Cas9 FOR EPIGENOME EDITING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
Cas9 FOR GENOMIC IMAGING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
Sequence-Specific Genomic Imaging Tools Based on Nucleotide
Base-Pairing Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
Sequence-Specific Genomic Imaging Tools Based on Protein–DNA Interactions . . 243
Cas9-Based Genomic Imaging: A Combination of Nucleotide Base-Pairing
and Protein–DNA Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
Challenges in Live Cell Genomic Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Cas9 FOR STUDYING ENDOGENOUS PROTEIN–GENOME
INTERACTIONS AT SPECIFIC LOCI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246

228 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

SPATIOTEMPORAL REGULATION OF Cas9 IN GENOMIC EDITING,


REGULATION, AND TARGETING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
Split Cas9s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
Cas9 FOR TARGETING RNA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
USE AND DELIVERY OF Cas9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Designing Single Guide RNAs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Choosing Target Sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Delivery Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
CONCLUSIONS AND FUTURE PERSPECTIVES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

INTRODUCTION: TOOLS FOR PROGRAMMABLE GENOME EDITING,


Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

TARGETING, AND REGULATION


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Since the advent of the central dogma of molecular biology, scientists have endeavored to develop
new technologies to modify or manipulate the genome. Precise editing and regulation of genomic
information is essential to understanding the function of a given gene. During the past decade,
technological breakthroughs have made genome editing and regulation significantly easier. One
recent technology has adapted the CRISPR (clustered regularly interspaced short palindromic
repeats)/Cas (CRISPR-associated protein) bacterial immune system as a simple, RNA-guided
platform for highly efficient and specific genome editing and regulation in diverse organisms,
thus creating revolutionary tools for biomedical research and new possibilities for treating genetic
disorders (1–14).
In general, the precise editing or regulation of genomic information at the DNA level requires
the action of a molecular machine composed of two major parts: a DNA-binding domain that
mediates sequence-specific DNA recognition and binding, and an effector domain that enables
DNA cleavage or regulates transcription near the binding site. Creating a double-stranded break
(DSB) by using a sequence-specific endonuclease can stimulate the DNA repair pathway and
greatly increase the rate of gene modification at the desired sequence (15–20). Thus, nuclease-
mediated approaches have been extensively explored for site-specific gene editing. Meganucleases,
or homing nucleases, are among the first classes of nucleases that were engineered to target specific
genomic sites for gene editing purposes (15, 16, 21). Meganucleases are a group of nucleases that
recognize long nucleotide sequences and induce a DSB at their targeted site. The long recognition
sequence of meganucleases may occur only once within a genome, thereby facilitating its use
for site-specific genome editing. Meganucleases can be reengineered to target novel sequences
through strategies such as protein engineering, structure-based design, and molecular evolution,
although the procedure is usually labor intensive (20–22).
Other examples of programmable genome editing machines include zinc-finger nucleases
(ZFNs) (23–25) and transcription activator-like effector nucleases (TALENs) (26–28), in which
the DNA-binding domains of transcription factors have been fused with the nuclease domain of
the restriction enzyme FokI, an obligate dimer (Figure 1a). When targeted to paired adjacent
sequences, the FokI domains of these programmable, site-specific nucleases form a dimer that ac-
tivates the nuclease activity, thus creating a DSB near their binding sites (Figure 1a). Researchers
can exploit the cell’s endogenous DNA repair pathways to create mutations at the desired DSB
sites. However, because these tools function through protein–DNA interactions, targeting to a
new site requires engineering and cloning a new protein, which precludes ZFNs and TALENs
from being used for high-throughput applications.

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 229


BI85CH10-Qi ARI 9 May 2016 9:29

In contrast to most known DNA-binding proteins, Cas9 is an RNA-guided nuclease whose


sequence specificity largely arises from Watson–Crick base pairing between its guide RNA and
the target DNA site, in addition to a direct interaction between Cas9 and a short protospacer-
adjacent motif (PAM) of DNA (3, 4, 13, 29, 30). Thus, Cas9 can be programmed to target new
sites simply by changing its guide RNA sequence, making it an ideal platform for high-throughput

a
Nucleases based on protein–DNA interactions

ZFN Zinc finger DNA-binding motif TALEN TALE DNA-binding motif

FokI FokI
FokI FokI
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

RNA-directed nucleases
Sp Cas9 Sa Cas9
Cas9
REC REC
REC sgRNA
3' sgRNA
20nt 5' HNH

Target DNA strand 3’


HNH 3’ sgRNA HNH WED
5' 3' WED
NUC
3' 5' DNA
RuvC DNA
Nontarget DNA strand
5’
PAM 5’
RuvC
Cleavage site RuvC PI PI

b Cas9
sgRNA

Nonhomologous end joining (NHEJ)


Mutations at targeted sites
Double-stranded break (DSB) (random insertions/deletions,
shift in reading frame, etc.)
nCas9 (nickase)

Silencing mutation
Homology-directed repair (HDR)
Double-stranded break (DSB)
Targeted sequence replacement
dCas9-FokI (mutations, gene correction,
gene knock-in, etc.)
Donor DNA
FokI
FokI

230 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

sequence-specific gene editing, as well as other applications. Its natural endonuclease activity has
been co-opted for sequence-specific editing of the genome in a wide range of organisms, including
bacteria (31), fungi (32), plants (33, 34), and animals (5–7, 9, 10, 35, 36). To enable sequence-
specific genomic regulation, nuclease-deactivated Cas9 (dCas9) has been engineered, and can
be fused to a variety of effectors, such as transcriptional activators, repressors, and epigenetic
modifiers (37–41).
In addition to applications in genome editing and regulation, DNA-binding proteins, such as
ZFs, TALEs, and dCas9, have been fused to fluorescent proteins (FPs) to allow direct imaging of
genomic loci in living cells (42–47). Additionally, dCas9 has also been used for studying proteins
that interact with specific loci (48, 49), and it may potentially be used to target RNA (50, 51). In
this review, we describe the working mechanism of Cas9 based on the findings of structural and
biochemical studies. We focus on the applications of CRISPR/Cas9 in genomic editing, regulation,
and imaging in mammalian cells, highlighting the power of this novel system in biological research.
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

CRISPR/CAS9: A GIFT FROM MOTHER NATURE

CRISPR: An Adaptive Immune Mechanism in Bacteria and Archaea


The CRISPR system is an adaptive immune mechanism present in many bacteria and the majority
of characterized Archaea. CRISPR-containing organisms acquire DNA fragments from invading
bacteriophages and plasmids before transcribing them into CRISPR RNAs (crRNAs) to guide
cleavage of invading RNA or DNA (1, 13, 29, 30, 52–56). This CRISPR immune system works
through the cooperation of many diverse Cas proteins. Based on differences in their components
and mechanisms of action, CRISPR systems have been divided into two major classes (57). RNA-
guided target cleavage in class 1 systems (types I, III, and IV) requires a large complex of several
effector proteins, but in the class 2 systems [type II, putative types V (58) and VI (59)], only
one RNA-guided endonuclease [e.g., Cas9 in type II and Cpf1 (CRISPR from Prevotella and

←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−

Figure 1
CRISPR-associated protein 9 (Cas9)-mediated sequence-specific genomic editing. (a) Comparison of programmable sequence-specific
genome editing nucleases. (Top) Zinc-finger nucleases (ZFNs) and transcriptional activator-like effector nucleases (TALENs) are
engineered by fusing ZF or TALE DNA-binding domains to the FokI nuclease domains (23–28). They recognize their targeted sites by
sequence-specific protein–DNA interactions, and a pair of ZFNs or TALENs cleaves adjacent sequences of the DNA to create a pair of
nicks on complementary strands, leading to a double-stranded break (DSB). (Bottom) Cas9 is a naturally evolved, RNA-guided nuclease.
It recognizes its target DNA through approximately 20 nucleotide (nt) base-pairing interaction between a single guide RNA (sgRNA)
and its targeted DNA strand. Cas9 also interacts with the protospacer-adjacent motif (PAM) of its DNA target through its
PAM-interacting (PI) domain at its C terminus. Cas9 uses its two nuclease domains (HNH and RuvC) to cleave the double-stranded
DNA, creating a DSB. The HNH, RuvC, and PI domains, as well as an evolutionarily divergent wedge domain (WED), all reside in
the Cas9 nuclease (NUC) lobe. The recognition (REC) lobe of Cas9 contains other regions that interact with the sgRNA–DNA
duplex. (Bottom right) Crystal structures of Sp Cas9 and Sa Cas9. Crystal structures of Streptococcus pyogenes Cas9 (Sp Cas9; Protein Data
Bank number 4UN3, 1368 AA) (84) and Staphylococcus aureus Cas9 (Sa Cas9 Protein Data Bank number 5CZZ, 1053 AA) (70) were
obtained from RCSB Protein Data Bank (http://www.rcsb.org/pdb/), compared using PyMOL (PyMOL Molecular Graphics System,
Version 1.3, Schrödinger LLC, https://www.pymol.org/), and domains are annotated according to References 70, 81, 83, 84. The
orientation of the target DNA strand is also shown. (b) Cas9 in genomic editing. The DSB generated by Cas9 activates the
nonhomologous end joining (NHEJ) or homology-directed (HDR) DNA repair pathways. NHEJ causes random insertions or
deletions (indels) at its targeted site, and HDR can create desired mutations or indels through homologous recombination guided by
donor DNA. A mutation in one nuclease domain of Cas9 creates a Cas9-based nickase (nCas9) that cleaves only one strand of DNA.
The specificity of Cas9-mediated genome editing can be greatly enhanced by using a pair of nCas9s that target each strand of DNA at
adjacent sites because both nCas9–sgRNA complexes must be present at the target site for DSB creation (5, 76–78). A similar strategy
has been achieved by using paired nuclease-deactivated Cas9 (dCas9)-FokI–sgRNA complexes (153, 154).

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 231


BI85CH10-Qi ARI 9 May 2016 9:29

Francisella-1) in type V] is required to mediate cleavage of invading genetic material (57–59).


Detailed descriptions of CRISPR system classification can be found in References 53, 54, 57, 59,
and 60.
In general, a CRISPR system works in three stages to carry out a full immune response to
invading foreign DNA (9, 13, 14, 53–56, 61, 62). In the first stage, or acquisition stage, DNA
fragments of invading plasmids or phages (termed protospacers) are incorporated into the host
CRISPR locus as spacers between crRNA repeats. In the second stage, Cas proteins are expressed,
the CRISPR array containing acquired spacers is transcribed into pre-crRNA, and the pre-crRNA
is cleaved and processed into mature crRNAs by Cas proteins and host factors (14). The fully
processed crRNA is a guide that contains a spacer sequence responsible for targeting it to the
invading genome, as well as all or part of the crRNA repeat sequence, which allows for recognition
of the crRNA by Cas proteins and other RNA components. In type II CRISPR systems, the
presence of a noncoding trans-activating CRISPR RNA (tracrRNA) that hybridizes with the
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

crRNA repeat sequence is critical for crRNA processing, Cas9 binding, and Cas9-mediated target
cleavage (3, 14). In the third stage, Cas proteins recognize the appropriate target with the guidance
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

of the crRNA and mediate the cleavage of the invading genome, thus protecting the host cells from
infection. The action of many CRISPR systems depends on the presence of a sequence-specific
PAM that is adjacent to the crRNA target site in the invading genome (30, 63–65). The absence
of this PAM sequence at the CRISPR locus in the host genome protects it from self-cleavage in
type I and type II CRISPR systems (9).

Repurposing Cas9 for Sequence-Specific Genomic Targeting


Many characterized Cas proteins bind to nucleic acids; thus, the CRISPR system can form the
basis of a flexible genomic engineering toolkit. Cas9, the RNA-guided endonuclease that cleaves
target DNA in the class 2 type II CRISPR system, is the most widely used for genomic editing
and regulation among the Cas proteins.
Cas9 target cleavage is guided by a duplex of two RNAs: the crRNA that recognizes the
invading DNA through an approximately 20–base pair (bp) Watson-Crick base-pairing region
and the tracrRNA that hybridizes with the crRNA and is unique to the type II CRISPR system
(3, 4, 12–14, 66). Cas9, in conjunction with the crRNA–tracrRNA duplex, can be repurposed
for efficient genomic editing (3, 5, 31). To simplify the system, a seminal work showed that
the crRNA–tracrRNA duplex can be fused into a chimeric single guide RNA (sgRNA) (3). This
single-protein, single-RNA, Cas9–sgRNA system is the most widely used for gene editing and
other Cas9-based applications.
The Cas9–sgRNA complex can bind DNA that base pairs with the sgRNA and is adjacent
to a PAM sequence (Figure 1a). Binding of the Cas9–sgRNA complex induces cleavage within
the base-pairing region. Thus, simply by customizing an approximately 20-nucleotide (nt) region
of the sgRNA to pair with the DNA sequence of interest, Cas9 can be retargeted to essentially
any genomic locus containing a PAM sequence, making it an easily programmable platform for
specific genomic targeting.

Diversity of Cas9 Orthologs


A large variety of Cas9 proteins exist in different bacterial type II CRISPR systems. These Cas9
nucleases range from about 900 to 1,600 amino acids (AA) in three subclasses: type II-A, type II-B,
and type II-C (9, 53, 54, 67). The most commonly used Cas9 for genome engineering has been
adapted from the type II-A CRISPR system from Streptococcus pyogenes (Sp). The Sp Cas9 has a

232 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

simple PAM (NGG, or a weaker NAG, where N is any nucleotide; 68) and has been optimized
for use in editing, as well as other contexts, using dCas9 across a wide variety of organisms.
Other Cas9 proteins have been studied and developed as tools. Of note is the Cas9 derived from
Staphylococcus aureus (Sa), which is 1,053 AA in length, with NNGRRT (where R is an A or G) as its
PAM (69). The relatively small size of Sa Cas9 allows it to circumvent some of the delivery issues
caused by the larger Cas9s, such as Sp Cas9 (1,368 AA; see the section Use and Delivery of Cas9
for further discussion). Sa Cas9 has been developed for genome editing (69) and gene regulation
(70) in mammalian cells, and it shows gene editing efficiency comparable to that of Sp Cas9 (69). In
addition to Sp Cas9 and Sa Cas9, other notable orthologs include Cas9 from Neisseria meningitidis
(Nm; PAM = NNNNGATT) and S. thermophilus 1 (St1; PAM = NNAGAAW, where W is an A
or T), which have been used in both bacteria and mammalian cells (71–74). Both Nm Cas9 and
St1 Cas9 have been engineered into dCas9 versions that have been used for gene regulation (38,
39, 71).
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

In addition to their distinct PAMs, these different Cas9s also have distinct crRNAs and tracr-
RNAs, which allow for the possibility of orthogonal genome editing, regulation, and imaging.
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Cas9–tracrRNA binding is sensitive to minor perturbations in the tracrRNA sequence and struc-
ture (75), which reinforces the orthogonality of these different Cas9s. Although there has been
some work using multiple Cas9s simultaneously to achieve distinct targeting and function (71, 72),
this area remains relatively underexplored.

Engineered Variants of the Cas9 Nuclease Domains: Nickase Cas9 (nCas9)


and Nuclease-Deactivated Cas9 (dCas9)
Cas9 contains two nuclease domains: an HNH nuclease domain that cleaves the target strand
of DNA (complementary to the guide RNA) and a RuvC-like nuclease domain that cleaves the
nontarget strand (Figure 1a) (3, 4). Mutating one of the two nuclease domains creates a nickase
Cas9 (nCas9) that cleaves only one strand of DNA (Figure 1b) (3, 4, 76–78). Two nCas9s can be
targeted to adjacent DNA sites to cause a DSB in a manner similar to that described for TALENs
above (77, 79). Pairs of nCas9s have been used to increase the specificity of Cas9-based genomic
editing, as only two adjacent nicking events will generate a DSB (76, 77, 80).
Mutating both nuclease domains generates dCas9, which lacks nuclease activity but retains its
RNA-guided DNA-binding activity. This allows dCas9 to be fused to other effectors to mediate
site-specific genetic and epigenetic regulation without cleaving the target DNA, as well as specific
DNA binding for several other applications (37–39, 41) (see discussion in sections Cas9 for Gene
Regulation, Cas9 for Epigenome Editing, and Cas9 for Genomic Imaging).

Other RNA-Guided Endonucleases in the CRISPR Systems


Recently, Zetsche et al. (58) discovered that, in the class 2 type V system, Cpf1 also mediates
RNA-guided target cleavage and can be adapted for gene editing in cultured human cells. In
contrast to Cas9, Cpf1-mediated DNA cleavage is guided by only a crRNA, and does not require
a tracrRNA. Additionally, Cpf1 uses different PAMs than those for characterized Cas9s, and it
creates a staggered DSB (58). Sequence analysis has revealed that Cpf1 contains only a RuvC-
like domain and lacks the HNH nuclease domain found in Cas9 (58). Based on the predicted
structure of effector proteins, Shmakov et al. (59) further classified three class 2 CRISPR systems,
including C2c1 and C2c3 (subtypes of putative type V) and C2c2 (a subtype of putative type
VI). They showed that C2c1 is a DNA endonuclease guided by both a crRNA and a tracrRNA.
The discovery of Cpf1 and other effector proteins in the diverse class 2 CRISPR systems further

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 233


BI85CH10-Qi ARI 9 May 2016 9:29

expands the toolkit of programmable RNA-guided endonucleases for genome editing (57–59).
These newly characterized proteins will no doubt be the source of many tool-building efforts in
the future.

STRUCTURE OF Cas9 AND PROPOSED WORKING MODEL FOR


GUIDE RNA BINDING AND TARGET DNA CLEAVAGE
The crystal structures of Cas9s have been reported, including Sp Cas9, Sa Cas9, and Cas9 from
Actinomyces naeslundii (Ana Cas9). The structure of Sp Cas9 has been characterized extensively: in
unbound (apo) form (81), sgRNA-bound form (82), and sgRNA–DNA-bound form (83, 84). The
Ana Cas9 structure has been resolved in its apo form (81), whereas the structure of Sa Cas9 has
been reported in complex with sgRNA and DNA (70). These studies have revealed the structural
details of interactions among Cas9, the sgRNA, and its DNA target, and also give insight into the
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

structural diversity of Cas9–sgRNA interactions.


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Bilobed Structure of Cas9


Cas9 adopts a bilobed architecture composed of a nuclease (NUC) lobe and an α-helical recog-
nition (REC) lobe (70, 81–83) (Figure 1a). The NUC lobe contains the HNH nuclease domain,
the RuvC-like nuclease domain, a PAM-interacting (PI) domain, and an evolutionarily divergent
wedge domain (WED) (70, 81, 83). The RuvC and HNH nuclease domains use, respectively, a
two-metal mechanism and a single-metal mechanism to cleave each of the DNA double strands
(70, 81, 83). The PI domain interacts with the PAM region of DNA through base-specific inter-
action and contributes to the DNA target specificity of Cas9 (70, 81, 83, 84). The WED domain
is important for orthogonal recognition of sgRNA scaffolds, and it also interacts with the back-
bone of the PAM region (70). The helical REC lobe is also diverse among different Cas9s, and it
contains regions that contribute to the recognition of guide RNA–target DNA heteroduplexes,
as well as specific recognition of cognate sgRNA scaffolds (70, 81, 83).
Both biochemical and structural studies have revealed that Sp Cas9 undergoes a series of
conformational changes to activate its DNA cleavage activity (70, 81–86); a working model is
detailed in the section Proposed Working Model for Guide RNA Binding and Target DNA
Cleavage.

DNA Targeting Specificity of Cas9: Protospacer-Adjacent


Motif (PAM) Interactions
The Cas9–sgRNA complex recognizes its DNA target through Watson–Crick base-pairing in-
teractions between the sgRNA and target DNA and through Cas9’s interactions with the PAM
adjacent to the sgRNA targeting site. The PI domain of Cas9 is composed of two domains: a
C-terminal domain and a topoisomerase-homology domain, adjacent to the C-terminal domain
(70, 84). Sp Cas9 recognizes a 5 -NGG-3 PAM, and its PI domain contains two arginine residues
(Arg1333 and Arg1335 ) that interact with the GG dinucleotides within the nontarget strand PAM
through base-specific hydrogen bonding (84). Sa Cas9 recognizes a 5 -NNGRRT-3 PAM, and
its PI domain contains several residues [e.g., asparagine (Asn)985 , Asn986 , Arg991 , and Arg1015 ] that
form base-specific hydrogen bonds with the GRRT bases (70). These base-dependent interactions
determine the specificity of PAM recognition by Cas9 orthologs.
Cas9 also contains a phosphate lock loop with residues [e.g., glutamate (Glu)1108 and serine
(Ser)1109 in Sp Cas9, and aspartate (Asp)786 and threonine (Thr)787 in Sa Cas9] that interact with the

234 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

target-strand DNA backbone directly adjacent to the PAM (70, 84). These interactions appear to
kink the DNA and facilitate the local, adenosine triphosphate–independent DNA strand separation
required to initiate the formation of the sgRNA–DNA duplex (70, 84).

DNA Targeting Specificity of Cas9: Guide RNA–Target DNA


Base-Pairing Interactions
Cas9–sgRNA targeting specificity is largely ensured by base-pairing interactions between the
sgRNA and its complementary target DNA strand. Mechanically, the activation of Cas9 nuclease
activity requires an HNH domain conformational change that depends on proper base-pairing
interactions between the guide RNA and its DNA target, providing another mechanism to ensure
the specificity of Cas9 in addition to PAM recognition and guide RNA–target DNA complemen-
tarity (86). The extent of this HNH domain conformational change, monitored by intramolecular
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Förster resonance energy transfer (FRET) assays, is sensitive to mismatches within the guide
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

RNA–target DNA base-pairing region (86).


Base pairing in the PAM-proximal region, referred to as the seed region, is where DNA double-
strand separation and sgRNA–DNA heteroduplex formation start. This has a crucial role in de-
termining the binding and cleavage specificity of Cas9 (3, 68, 69, 87–90). The PAM-distal regions
are more tolerant of mismatches as assayed by Cas9 binding and cleavage (3, 68, 69, 87). This is
consistent with the model in which the Cas9–sgRNA complex first surveys the genome for the
PAM site before unwinding DNA, starting with the PAM-proximal portion of the target DNA
sequence (87).
Structurally, the PAM-proximal RNA seed region bound to Cas9 maintains an A-form confor-
mation to facilitate sgRNA–DNA heteroduplex formation (70, 82). Biochemical work has revealed
that binding to at least 10 nt of the guide RNA seed region is required to trigger a conformational
change within Sp Cas9 for DNA recognition (82), and FRET assays have shown that shorter guide
RNAs induce lower levels of conformational changes than a full-length guide RNA (86).

Interactions between the Single Guide RNA–DNA Duplex and Cas9


Structural studies also have demonstrated that the sgRNA–DNA duplex interacts with Cas9
through sequence-specific and nonspecific interactions, providing more insights into its func-
tional mechanism (70, 82–85). The sgRNA consists of three key regions from 5 to 3 : a guide
RNA–target DNA heteroduplex region where the sgRNA base pairs with the target DNA, a
repeat–antirepeat duplex that represents a hybridization region between crRNA and tracrRNA,
and additional stem-loops that are found in the tracrRNA in the endogenous CRISPR locus (3, 14,
83). In general, the guide RNA–target DNA heteroduplex and the repeat–antirepeat duplex are
both located in a positively charged groove between the two lobes of Cas9. The additional sgRNA
stem-loops also interact with charged residues on the surface of Cas9 to enforce the interaction
between Cas9 and its cognate sgRNA.
Cas9 recognizes the guide RNA–target DNA heteroduplex region in a sequence-independent
manner, primarily through interactions with its phosphate backbone, but interactions between
many other regions of the sgRNA and Cas9 depend on the sgRNA sequence and folding (70, 75, 83,
85). These aspects of Cas9 structure allow it to flexibly target any PAM-adjacent DNA sequences
with paired sgRNAs while at the same time precisely recognizing guide RNAs containing specific
sequences and structures (70, 83, 85).

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 235


BI85CH10-Qi ARI 9 May 2016 9:29

Orthogonality of Cas9–Single Guide RNA Interactions


The comparison of different Cas9 structures provides insights into the orthogonal DNA-targeting
mechanism of Cas9s (70, 81, 83, 84). In addition to binding to their specific PAM motifs, Cas9 or-
thologs recognize their cognate sgRNA scaffolds through sequence-specific and structure-specific
interactions (70, 83, 84). For example, the repeat–antirepeat duplexes are significantly different
between Sp sgRNA and Sa sgRNA. Their distinct structural features are recognized by the WED
domains and REC lobes of their respective Cas9s in a highly specific manner (70, 83, 84). More-
over, three stem-loops of the Sp sgRNA are required for efficient Cas9-mediated DNA cleavage
in vivo (68, 83), but sequence predictions suggest that the Sa sgRNA may contain only two stem-
loops (69, 70). The phosphate backbones and some residues in these stem-loops also interact with
different regions of their cognate Cas9s in a structure-specific or base-specific manner (70, 83, 84).
The structural dependence of Cas9–sgRNA interactions forms a basis for orthogonal recognition
of sgRNA and, thus, orthogonal DNA targeting, and it underscores the importance of maintaining
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

interaction-relevant nucleotides when optimizing the sgRNA scaffold.


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Proposed Working Model for Guide RNA Binding and Target DNA Cleavage
A working mechanism of Cas9 has been proposed by combining structural studies (70, 81–85) and
in vitro assays (86–88, 91). In this model, the Cas9 protein maintains an autoinhibited conformation
when not bound by sgRNA in which the active sites in the HNH domain are blocked by the RuvC
domain (81). The binding of an sgRNA induces a conformational change to create a central channel
between the two lobes for DNA binding (70, 81–86), thus entering into a DNA recognition–
competent state (82). The resulting Cas9–sgRNA pretargeting complex can survey DNA for PAMs
by three-dimensional diffusion (87, 92). The Cas9–sgRNA complex binds to a PAM through its
PI domain, which initiates local DNA strand separation in the PAM-proximal region to facilitate
sgRNA–DNA heteroduplex formation (84). The Cas9–sgRNA complex will continue to unwind
the DNA only if there is a significant match between the guide RNA segment and the target
DNA (82, 86). The strong guide RNA–target DNA base-pairing interactions further promote
DNA double strand separation and RNA–DNA heteroduplex formation, which proceeds from
the PAM-proximal region and forms a complete R loop (86–88). Finally, the complete R loop
causes another conformational change in the HNH domain, activating the nuclease activity of
both the RuvC and HNH domains to induce DNA cleavage (81–88). Sp Cas9 creates a DSB
approximately 3 nt away from the PAM in the target DNA (3,4).

Cas9 FOR GENOME EDITING

Mechanism of Genome Editing: DNA Cleavage Followed by DNA Repair


Since its discovery, Cas9 has been extensively used for genome editing in multiple organisms.
Cas9, like engineered ZFNs and TALENs, is a programmable, sequence-specific endonuclease.
Similar to other nucleases, Cas9-mediated genome editing is achieved by a two-step process: DNA
cleavage followed by DNA repair (Figure 1b). The sgRNA directs Cas9 to a specific genomic
locus where Cas9 creates a DSB (3, 4), which triggers DNA repair through intrinsic cellular
mechanisms, such as nonhomologous end joining (NHEJ) or homology-directed repair (HDR)
(15–19).
NHEJ causes nearly random insertion and deletion mutations (i.e., indels) at the DSB site
and, thus, may lead to gene knockout (e.g., by causing a shift in the target gene’s reading frame

236 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

or mutating a critical region of the encoded protein) (Figure 1b) (93). HDR can be exploited to
generate the desired sequence replacement at the DSB site through homologous recombination
guided by a donor DNA template, causing targeted gene deletion, mutagenesis, insertion, or gene
correction (Figure 1b) (17, 19). Thus, the CRISPR/Cas9 system provides a powerful platform
for sequence-specific genome editing, including gene knockout, gene knockin, and site-specific
sequence mutagenesis and corrections (9, 10, 35).

Applications of Cas9-Mediated Genome Editing for Studying Gene Function,


Disease Modeling, and Gene Therapy
The Cas9-mediated gene-editing system has been broadly used in reverse genetics studies to
understand the role of specific genes, for disease modeling, and for demonstrating new therapeutic
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

schemes in a number of models of genetic and infectious diseases (9, 10, 94).
Retargeting Cas9 to a new DNA site is easy to achieve by simply creating a new sgRNA that
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

pairs with the desired DNA targeting site adjacent to PAM. In the case of Sp Cas9, the NGG
PAM motif occurs, on average, once every 8 bp within the genome, thus allowing almost any gene
of interest to be targeted (9, 10, 35). Cas9s from other species have different PAMs, of different
sizes and comprising a variety of sequences, which further expands the range of Cas9-targetable
genomic sequence [e.g., Sa Cas9 (69), St1 Cas9, and Nm Cas9 (71)]. The engineering of existing
Cas9s has also led to the creation of new versions of Cas9 with altered PAM sequences (95, 96),
thus expanding the targetable space within the mammalian genome.
The use of the Cas9 platform has greatly increased the efficiency of generating transgenic
organisms, from fungi (32) and plants (33, 34, 97, 98) to a variety of animals (5–7, 36, 99–102)
(reviewed in 9, 10, 35). This technology also makes it much easier to generate disease models
for genetic disorders and diseases such as cancer, which aids our understanding of the molecular
mechanisms of these pathological processes (reviewed in 9, 10, 103).
Cas9 can be easily programmed to edit multiple genomic loci at the same time by introduc-
ing several sgRNAs simultaneously. This can be applied to generate large-scale chromosomal
rearrangements. For example, creating a pair of DSBs at nearby regions within the same chro-
mosome may produce targeted deletions or inversions of the intermediate segment of DNA
(104–110), and creating two DSBs in different chromosomes may lead to a targeted chromosomal
translocation (107, 111). These Cas9-mediated, targeted rearrangements may be useful for creat-
ing disease models by mimicking rearrangements that occur in human disease states (e.g., cancers
and heritable genetic disorders) (107, 110, 111).
The Cas9 system also has the potential to cure or treat many maladies, including HIV, genetic
diseases, and cancer (94, 112). For example, when Cas9 is introduced into infected cells together
with sgRNAs targeting crucial viral genomic elements, it helps to inactivate or clear the viral
genome and, thereby, defends the cells or organism from infections with HIV (113, 114), hepatitis
B virus (115–117), human papillomavirus (118), and Epstein–Barr virus (119). Moreover, it has
been shown by using CRISPR/Cas9 (120, 121) or ZFNs (122–124) that editing the genes of HIV
coreceptors (e.g., CCR5) in the host genome, which encodes a coreceptor of HIV, creates cellular
resistance to the HIV-1 virus and, thereby, may help to combat infection.
In addition, many studies have reported using the Cas9-mediated genome editing system for
correcting disease-related mutations in animal somatic (125) and germ line cells (126–128), as
well as in human stem cells (129) and induced pluripotent stem cells (130–136). A partial list
includes the Fah gene in hereditary tyrosinemia (125), Dystrophin in Duchenne muscular dystrophy
(126, 133), Crygc in cataracts (127, 128), CFTR in cystic fibrosis (129), HBB in β-thalassemia

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 237


BI85CH10-Qi ARI 9 May 2016 9:29

(132, 134, 135), JAK2 in polycythemia vera (136), and SERPINA1 in α-1 antitrypsin deficiency
(136) (reviewed in 94, 112, 137).

Applications of Cas9-Mediated Genome Editing for Genome-Wide


Functional Screening
Significantly, the Cas9 platform has been used for large-scale genome-wide knockout screens
that had been previously unfeasible (138–145). Previously, genomic loss-of-function screening
relied on the RNA interference (RNAi) approach, which represses gene expression at the RNA
level without affecting the DNA sequence (146–148). In RNAi, a small interfering RNA (siRNA)
that base pairs with its target messenger RNA (mRNA) will lead to a decrease in the stability and
translation of its target. The siRNA can be synthesized or produced from a vector encoding a short
hairpin RNA (shRNA), an artificial RNA molecule containing a hairpin that is then processed into
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

the mature siRNA form by the cell’s endogenous small RNA pathway. In this way, large-scale
gene knockdown screening can be achieved using a library of siRNAs or shRNAs.
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Similarly, by creating a library of sgRNAs targeting gene coding regions, researchers can exploit
the CRISPR/Cas9 platform to screen for genes contributing to a biological process. The Cas9–
sgRNA approach generates indels at the targeted loci and may cause complete loss of gene function,
whereas the RNAi method may lead to only partial gene suppression. Thus, when targeting the
same gene, the CRISPR/Cas9 technique may generate a more pronounced phenotype than RNAi,
which may make identification of relevant genes easier. One avenue to validate candidate genes
identified with the CRISPR/Cas9 approach is to re-express the targeted gene in the knockout
strain (143). Similarly, hits discovered with the RNAi approach may be validated by expression of
an RNAi-resistant transcript (143).
In terms of limitations in targeting, the CRISPR/Cas9 method can target only a sequence ad-
jacent to PAM, and not all exons contain such a targetable sequence, whereas an siRNA or shRNA
library, in principle, can target any mRNA sequence. In addition, a complete gene knockout by
CRISPR/Cas9 requires all alleles of the same gene to be mutated, which makes the screening
more challenging for cells containing several alleles, such as cancer cells (147). Furthermore, use
of the CRISPR/Cas9 knockout approach to study essential genes is challenging, because deletion
of essential genes causes a lethal effect that prevents most functional assays. Both methods can form
the basis of a successful screen, and the method choice will depend on the needs of the experiment.

Challenges in Cas9-Mediated Genome Editing


Despite Cas9’s great potential for both research and therapeutics, improvements can still be made
in its specificity, efficiency, and spatiotemporal control (149). One concern with the commonly
used Sp Cas9 system is the possibility of off-target effects because the 20-bp targeting sequence in
the sgRNA plus the 3 bp PAM may potentially be present elsewhere in the genome (68, 150, 151).
Different strategies have been developed to improve Sp Cas9 specificity: optimizing the sgRNA
design (68, 142, 152); using paired nCas9s (5, 76–78), paired dCas9-FokI nucleases (153, 154)
(Figure 1b), enhanced Sp Cas9 with improved specificity (155), shorter (17–18 bp) sgRNAs or
sgRNAs with two unpaired Gs on the 5 end that are more sensitive to mismatches (156, 157); or
decreasing the concentration of the Cas9–sgRNA complex or its length of active time within the
cell (68, 158). Although these strategies have greatly improved specificity, they sometimes come
at the cost of efficiency.
Another major challenge is to improve the efficiency of precise genome editing via HDR while
reducing indel generation through NHEJ. To address this, strategies have been developed to

238 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

modulate the HDR:NHEJ ratio, including altering the expression of DNA repair components,
using small molecules, synchronizing the cell cycle, and optimizing delivery timing and methods
(159–163). It is also imperative to develop tightly regulated platforms, as well as safe and efficient
delivery methods, for precise control of Cas9 activity, especially for potential applications in gene
therapy.

Responsible Use of Cas9


The rapid progress of Cas9 as a tool for genome editing has transformative potential for use in
applications ranging from clinical treatments to agricultural production to population control of
disease-carrying insect species. The rapid advances in Cas9 technology have introduced challenges
concerning the regulatory controls governing its safe, secure, and ethical applications. Concerns
within the community flared up after it was reported that one group had used Cas9 to edit a
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

gene in human embryos, although this was done in nonviable, triploid zygotes (164). There is
much debate among scientists, bioethicists, policymakers, and the public about how to ethically
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

and responsibly use gene editing technology in a way that does not hamper beneficial scientific
research and discovery (165–171). Regarding the editing of human cells, major questions include,
but are not limited to, should editing of human somatic tissues or germ-line cells be allowed and
regulated? Should this technology be used in human embryos? If the answer to these questions is
yes, in which cases would this be applicable?
In addition to the ability to edit the human genome, Cas9 also offers the possibility of drasti-
cally altering ecosystems by editing the genomes of plants and animals. By editing the genomes
of crops and livestock, there is a potential to greatly increase the yield of food production. Cas9-
based genome editing technology also has been proposed as a possible method for controlling the
populations of disease transmitters, such as mosquitoes that transmit malaria (172). This could be
accomplished through the use of gene drive technology, which facilitates the rapid spread of ge-
nomic alterations in wild populations (reviewed in 173). Although there is potential benefit to using
Cas9-based gene drives, there is still much debate about how or if this technology should be used.
Major concerns include doubts about our ability to predict the full ecological impact of such genet-
ically modified organisms and our ability to contain or control them once released into the wild.
The rapid development of Cas9 technology underscores our need as a scientific community
and as a society for a comprehensive policy regarding the use of genome editing technology. As the
rapid pace of biological discovery continues, discussion will be necessary to ensure that genome
editing technologies, such as Cas9, will be used in a safe and responsible manner.

Cas9 FOR GENE REGULATION: CRISPR INTERFERENCE (CRISPRi)


AND CRISPR ACTIVATION (CRISPRa)

Nuclease-Deactivated Cas9 (dCas9): A Programmable Platform for


Sequence-Specific Gene Regulation
In addition to its nuclease activity, Cas9 can serve as a unique platform to recruit protein and RNA
factors to a targeted DNA site, and it has been engineered into powerful tools for sequence-specific
gene regulation (37, 41, 174, 175). To achieve this, transcriptional activators and repressors are
fused to dCas9; dCas9 maintains its ability to bind both the sgRNA and targeted DNA, but it lacks
nuclease activity and, thus, can serve as a sequence-specific RNA-guided DNA-binding platform.
In bacterial cells, dCas9 alone can efficiently inhibit the transcription of targeted genes through
steric hindrance of transcriptional machinery (Figure 2a) (41, 176). This novel technique is

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 239


BI85CH10-Qi ARI 9 May 2016 9:29

termed CRISPR interference (CRISPRi), as it interferes with the transcription of RNA. Although
CRISPRi is generally highly efficient in prokaryotes, the dCas9–sgRNA complex alone may not
be very efficient at silencing gene expression in mammalian cells (41). However, CRISPRi in
mammalian cells can be enhanced by fusing dCas9 to a transcriptional repressor domain (e.g., the
KRAB domain of Kox1), which leads to successful suppression of reporter and endogenous genes
(Figure 2a) (37, 177).
In addition to CRISPRi, CRISPR activation (CRISPRa) has been created by fusing dCas9
to transcriptional activators, such as VP64 and p65AD in mammalian cells (Figure 2b) (37, 76,
175, 177) and the ω subunit of RNA polymerase in bacteria (176). These dCas9 fusions are
able to upregulate gene expression in host cells. In this review, we focus on the use of CRISPRi
and CRISPRa (CRISPRi/a) in eukaryotic cells. CRISPR/Cas9-based prokaryotic activation and
repression is reviewed in Reference 178.
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Multimodal CRISPRi/a Function


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

In addition to direct fusions of an activator or repressor to dCas9, the sgRNA can be modified and
turned into a scaffold to recruit transcriptional regulators (76, 179–181). The sgRNAs can be fused
to orthogonal protein-interacting RNA aptamers, which recruit specific RNA-binding proteins
(RBPs) (Figure 2c). These aptamer-modified sgRNAs are termed scaffold RNAs (scRNAs) (179).
Transcriptional activators and repressors can be fused to these RBPs in lieu of dCas9. When
orthogonal RNA aptamer–RBP pairs (e.g., MS2–MCP, PP7–PCP, com–Com) are coupled to
different sgRNAs, distinct RBP transcriptional modules can be recruited to different genes to
achieve multimodal regulation (i.e., simultaneous activation and repression) (179). For example,
in the presence of Sp dCas9, one gene can be targeted by an scRNA with an aptamer that will
recruit VP64 and cause activation. Another gene can simultaneously be targeted by an scRNA
with an aptamer that will recruit KRAB and cause repression (Figure 2c) (179). Thus, this system
allows for multimodal regulation of different genes within the same cell using a single Sp dCas9
protein (179).

Strategies to Improve CRISPRa Efficiency


The efficiency of CRISPRa can be dramatically enhanced by recruiting multiple transcriptional
activators to upregulate gene transcription. In addition to using multiple sgRNAs tiled along the
promoter to recruit multiple dCas9 activators (182–184), other strategies have been developed to
recruit multiple transcriptional activators to a dCas9-binding site (180, 184, 185). For example,
the synergistic activation mediator (SAM) system uses both dCas9 and sgRNA as scaffolds to
recruit multiple activators that function synergistically to enhance the activation of endogenous
genes (180). In this system, dCas9-VP64 is combined with a modified sgRNA containing two
MS2 RNA aptamers. Each MS2 aptamer recruits a pair of cognate RNA-binding proteins, MS2
bacteriophage coat proteins (MCPs), which are fused with the activating domains of p65 and HSF1
(MCP-p65-HSF1) (Figure 2b) (70, 180). This system increases activation efficiency and has been
applied to large-scale genomic screening (180).
Another strategy to enhance activation was developed by Tanenbaum et al. (185) and Gilbert
et al. (40). In these two articles, the authors combined the dCas9 system with a recently developed
multipeptide array, SunTag, to recruit multiple VP64 activator modules to a single dCas9-binding
site. Specifically, dCas9 was fused to an array of polypeptides (GCN4s) that can recruit multiple
copies (e.g., 10 or 24 copies) of its cognate single-chain variable fragment (scFv, an engineered
portion of an anti-GCN4 antibody). The scFv was then fused to VP64, leading to the recruitment

240 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

a Gene repression (CRISPRi) b Gene activation (CRISPRa) HSF1

VP64 p65
dCas9
MCP
sgRNA
MS2

RNAP

dCas9-VP64
dCas9 alone (bacteria)
dCas9 SAM system

GCN4 p65 Rta


VP64
ScFv-VP64
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

dCas9-KRAB dCas9 SunTag system dCas9-VPR

c Orthogonal gene repression and activation with scRNA


RNA aptamer (MS2) VP64 com KRAB
RBP (MCP) RBP (Com)

scRNA

Gene 1 Gene 2

d Epigenetic modification

dCas9
sgRNA ac
me

Gene Gene

dCas9-LSD1 dCas9-p300 core

Figure 2
Nuclease-deactivated Cas9 (dCas9)-mediated sequence-specific gene regulation. (a) CRISPR interference (CRISPRi) strategies.
Repression can occur with dCas9 alone in bacteria, which sterically blocks transcriptional elongation of RNA polymerases (RNAPs)
(41). Alternatively, dCas9 can be fused to a repressor domain such as KRAB to enhance repression (37). (b) CRISPR activation
(CRISPRa) strategies. Activation can be achieved by directly fusing dCas9 to a transcriptional activator (e.g., VP64) or by recruiting
multiple transcriptional activators using the synergistic activation mediator (SAM), SunTag, or VP64-p65-Rta (VPR) systems (37, 180,
184, 185). Note for the SAM system, each MS2 aptamer can recruit a pair of MCP-p65-HSF1, but only one is shown for simplicity.
(c) Gene activation and repression can occur simultaneously in the same cell using the scaffold RNA (scRNA) system. RNA aptamers
(e.g., MS2, com, PP7) are fused to single guide RNA (sgRNA), creating an scRNA that is localized to a specific genomic locus with
dCas9. The scRNAs can recruit RNA-binding proteins (RBPs; e.g., MCP, Com, PCP) fused to an activator (e.g., VP64, left) or a
repressor (e.g., KRAB, right) (179). (d ) CRISPR-mediated epigenetic modification. The epigenetic landscape can be altered in a
site-specific manner by fusing epigenetic modifying enzymes such as p300 or LSD1 to dCas9. For example, dCas9-LSD1 decreases
H3K4me2 near the targeted enhancer region, resulting in repression of related genes, whereas dCas9-p300 increases H3K27
acetylation at the promoter or enhancer regions and activates the expression of downstream genes (38, 39).
www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 241
BI85CH10-Qi ARI 9 May 2016 9:29

of multiple copies of VP64 to each dCas9 (Figure 2b) (185). This system has been used to
strongly upregulate chemokine (C-X-C motif) receptor 4 (known as CXCR4), thus enhancing
cell migration in K562 cells. Additionally, Tanenbaum et al. (185) upregulated cyclin-dependent
kinase inhibitor 1B (known as CDKN1B) using the SunTag system, leading to a reduction in cell
growth. The system has also been used for genome-scale gain-of-function screening (40).
In a third strategy, used by Chavez et al. (184), dCas9 was fused to three different activators in
tandem, VP64-p65-Rta (VPR), resulting in a tripartite activator. dCas9-VPR was able to achieve
greater activation of endogenous genes than dCas9-VP64 (Figure 2b). This system has been
used to direct the neuronal differentiation of induced pluripotent stem cells with multiple coex-
pressed sgRNAs (184). All of these studies show that synergistically recruiting multiple activators
to the dCas9 target locus enhances the activation of the CRISPRa system (180, 184, 185). These
engineered systems likely mimic intrinsic cellular gene activation mechanisms, which work by
coordinating the recruitment of multiple activators (186, 187).
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Advantages of CRISPRi/a and Their Applications


Compared with RNAi and other established methods of regulating gene expression (e.g., gene
overexpression, TALE- or ZF-mediated regulations), CRISPRi/a combine the advantages of de-
sign simplicity (37, 174, 188) with high specificity (37, 175, 180, 189), directly control endogenous
gene expression at the transcriptional level, and can act on both coding and noncoding sequences
(40, 180, 184, 190) (reviewed in 191, 192). CRISPRi/a have been applied in large-scale genomic
screening, showing minimal off-target effects in different systems (40, 180). The high specificity
of CRISPRi/a may be attributed to their narrow regions of activity (i.e., around the transcriptional
start site, or TSS; discussed in the section Use and Delivery of Cas9) and their high sensitivity
to sgRNA–DNA mismatches (40). In addition, CRISPRi/a have also been used to regulate the
transcription of non-protein coding RNAs, such as long noncoding RNAs (40, 180) and micro
RNAs (190). Other work has also expanded the applications of CRISPRi/a to the regulation of
gene expression in multicellular organisms, such as Drosophila (193) and plants (194, 195), and to
the reactivation of latent reservoirs of HIV-1 for its permanent elimination (196).

Cas9 FOR EPIGENOME EDITING


In a manner similar to that used to regulate transcription, sequence-specific DNA-binding pro-
teins can recruit epigenetic modifiers to reshape the epigenome at a given locus. It has been
shown that ZFs and TALEs fused with epigenetic modifiers can alter epigenetic marks at their
target DNA sites, which can lead to changes in relevant gene expression (197–202). Some studies
have also reported the use of dCas9 systems to achieve site-specific epigenome editing (38, 39,
203). Kearns et al. (39) fused Nm dCas9 with the histone demethylase LSD1 (Figure 2d ). They
then targeted enhancers of genes (e.g., Oct4, Tbx3) that are crucial for maintaining pluripotency
in mouse embryonic stem cells (mESCs). They demonstrated that Nm dCas9-LSD1 efficiently
suppressed the expression of genes controlled by the targeted enhancers, decreased the level of
the epigenetic marks H3K4me2 and H3K27ac near the targeted Tbx3 enhancer region, and also
caused changes in cell morphology. Another study revealed that the dCas9-KRAB fusion can
induce H3K9 trimethylation (H3K9me3) when targeted to the HS2 enhancer and suppress the
expression of globin genes that is regulated by the HS2 enhancers (203).
Using a slightly different approach than those described above, Hilton et al. (38) fused Sp dCas9
and Nm dCas9 to the catalytic core domain of the histone acetyltransferase p300 (Sp dCas9-p300
core and Nm dCas9-p300 core, respectively) (Figure 2d ). The authors were able to activate the

242 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

expression of several endogenous genes by targeting the promoter or enhancer regions, and they
showed that the activation was specific using genome-wide RNA sequencing. When targeted to
the HS2 enhancer, dCas9-p300 core increased the level of H3K27 acetylation at both the targeted
enhancer and the promoters of its downstream genes.
The studies (38, 39, 203) described above have demonstrated that dCas9 fusion proteins can
act as sequence-specific, synthetic epigenome modifiers, which not only change local epigenetic
status but also change the gene expression of relevant genes. Given the broad expanse of functional
epigenetic marks—from DNA methylation to histone modifications—future studies are needed
to develop a full toolkit of dCas9-based epigenetic modifiers. Given the possible off-target effects
of dCas9 when using dCas9-mediated epigenome editing systems to target a specific locus, the
specificity and toxicity of such tools should also be assessed.

Cas9 FOR GENOMIC IMAGING


Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

In the postgenomic era, another challenge for scientists is to understand the correlations between
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

the linear genetic information within DNA and its three-dimensional organization within the cell
nucleus. Many studies have revealed that the three-dimensional organization of genomic structure
may play an important part in regulating gene expression and controlling cell differentiation (204–
207). Further research into the correlations between genomic architecture, gene expression, and
cell behavior is hindered by the lack of tools for visualizing sequence-specific genomic dynamics
in living cells. Cas9’s ability to localize to specific sequences within the genome and the ease of
redirecting it to different genomic loci have made it a promising candidate for studying genomic
organization and dynamics in living cells.

Sequence-Specific Genomic Imaging Tools Based on Nucleotide


Base-Pairing Interactions
In principle, the labeling of a specific genomic locus in the nucleus can be achieved through either
nucleotide base-pairing interactions or sequence-specific protein–DNA interactions (Table 1).
Labeling techniques that rely on nucleic acid interactions, such as in situ hybridization (ISH)
assays, have been developed and used extensively in genomic research and for the clinical diag-
nosis of genetic diseases. ISH uses in vitro synthesized and labeled nucleotide probes to visualize
complementary endogenous genomic loci (208–210). For example, multicolor, fluorescent in situ
hybridization (FISH) uses fluorescently labeled nucleotide probes to simultaneously detect the lo-
calization of multiple loci (211–215). Likewise, electron microscopy in situ hybridization (known
as EM-ISH) uses radioactively labeled probes or biotin and digoxigenin labeled probes to detect
genomic ultrastructure (216, 217).
Although powerful, these ISH assays are generally restricted to fixed samples because cells
need to be fixed, permeabilized, and their DNA denatured before the labeled probes can bind.
One exception to this, reported by Molenaar et al. (218), demonstrated that a peptide nucleic acid
(PNA) probe can be introduced into the cells by glass beads to track the dynamics of telomeres in
living cells. Whether a similar system can be used to label other genomic loci remains untested.

Sequence-Specific Genomic Imaging Tools Based on Protein–DNA Interactions


An alternative to ISH techniques is the labeling of genomic loci through sequence-specific
protein–DNA interactions. This technique is feasible for live cell imaging, given the ease of
fusing DNA-binding proteins with FPs and expressing these constructs in living cells. Moreover,

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 243


Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.
BI85CH10-Qi

Table 1 Comparison of sequence-specific genomic imaging methods: Nucleic acid probes (left), imaging based on protein–DNA interactions (middle), and
CRISPR/dCas9-based genomic imaging (right) have different advantages and disadvantages based on the given application
ARI

Combined protein–DNA and


nucleotide base-pairing
Mechanisms Nucleotide base-pairing interactions Protein–DNA interaction interactions

244
Labeled nucleotide probe Fluorescent protein (FP) Fluorescent protein (FP)
9 May 2016

dCas9

Wang
sgRNA
Genome

·
9:29

Sequence-specific DNA binding proteins

La Russa
·
PNA probe live Native DNA-binding

Qi
Technique ISH imaging LacO/TetO proteins ZF/TALE CRISPR/Cas9
Probe Labeled nucleic acids Labeled PNA probe Lac/Tet repressors-FP Endogenous ZF/TALE-FP dCas9-FPs + sgRNAs
strand (DNA, RNA, DNA-binding
and PNA, etc.) protein-FP
Genomic Any genomic loci Telomeres Genomically integrated Genomic loci containing Repetitive genomic loci Essentially any genomic locus
target LacO or TetO arrays repetitive native (repetitive or nonrepetitive
protein-binding sites sequences)
(telomeres, etc.)
Cells Fixed cells Living cells Living and fixed cells Living or fixed cells (i.e.,
CASFISH)
Advantages  Easy probe  Enables live  Enables live cell imaging  Enables live cell imaging
design cell imaging  Allows multiple color imaging  Easy sgRNA probe design
 Allows multiple  Multiple color imaging with
color and orthogonal dCas9s or
high-resolution CASFISH
imaging
Disadvantages  Restricted to  Only shown for  Laborious to  Restricted to genomic  Laborious to  Requires multiple sgRNAs to
fixed samples telomere live create and loci that have natural construct many image a nonrepetitive
 Lacks dynamic cell imaging characterize sequence-specific ZF/TALE sequence
information  Challenges in insertions binding proteins proteins  Requires a PAM
delivery  Cannot directly  So far only
label endogenous restricted to
genomic loci repetitive
sequences
References 208–217 218 219–221 222–224 42–46 47, 72, 185, 225, 226

Abbreviations: Cas9, CRISPR-associated protein 9; FP, fluorescent protein; ISH, in situ hybridization; PAM, protospacer-adjacent motif; PNA, peptide nucleic acid; sgRNA, single guide RNA; TALE, transcription
activator-like effector; ZF, zinc finger.
BI85CH10-Qi ARI 9 May 2016 9:29

some DNA-binding proteins have high binding specificity for cognate DNA sequences in living
cells, and these protein–DNA interactions do not require DNA denaturation.
Initial work to visualize genomic dynamics involved the insertion of Lac/Tet operator tandem
repeats into a specific genomic locus. These exogenously added repeats were then visualized
using their binding proteins fused with FPs (219–221). This method allows us to understand the
dynamics of genomic loci in living cells. However, it is labor intensive to insert repetitive tandem
repeat sequences into a specific genomic locus, and this method cannot be used to directly label
endogenous sequences.
One way to visualize endogenous loci is to co-opt endogenous DNA-binding proteins and
label them with FPs. Some repetitive genomic loci, such as telomeres and centromeres, have na-
tive sequence-specific binding proteins and, thus, can be easily visualized by fusing these binding
proteins to FPs or by immunostaining with related antibodies (222–224). However, the majority
of the human genome sequence lacks unique binding proteins. For this reason, the use of pro-
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

grammable DNA-binding proteins, such as ZFs, TALEs, and Cas9s, offers a powerful approach
for imaging these genomic loci.
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Several groups have used ZFs or TALEs to image endogenous genomic loci. Initial efforts were
made by fusing green fluorescent protein (GFP) to ZF proteins to image repetitive sequences
at pericentric regions in living cells (42). Several studies (43–46) have also imaged repetitive
genomic elements by fusing FPs to TALEs. Miyanari et al. (43) reported a TALE-mediated
genome visualization method used to track the dynamics of centromeres and telomeres in living
mESCs and mouse embryos, and they also used this method to efficiently distinguish two parental
chromosomes with distinct single nucleotide polymorphisms, suggesting a high specificity. Ma
et al. (44) published similar and complementary results showing that two TALEs tagged with
different colors can simultaneously track the dynamics of centrosomes and telomeres in living
cells. They also showed in vitro purified TALE proteins could label genomic loci in fixed cells by
using a protocol simpler than FISH. Another work by Thanisch et al. (45) used a FP-TALE to
track the dynamics of satellite repeats throughout the cell cycle in mESCs.

Cas9-Based Genomic Imaging: A Combination of Nucleotide Base-Pairing


and Protein–DNA Interactions
Cas9-based imaging approaches combine the advantages of nucleotide interactions and protein–
DNA interactions to label endogenous genomic loci (Table 1). Soon after TALE-mediated
genomic imaging was reported, the first exciting work using dCas9 for genomic imaging was
published (47). In this work, Chen et al. (47) fused Sp dCas9 to enhanced GFP to visualize the
dynamics of the genomic loci of coding and noncoding sequences in living human cells. In this
study, repetitive genomic loci were dynamically tracked throughout the cell cycle using a single
sgRNA. A nonrepetitive genomic locus can also be labeled by co-delivering multiple sgRNAs
that tile the locus (47). Another group used a similar strategy to label endogenous centromeres,
pericentric regions, and telomeres in living mESCs (225).
There have been several additions made to improve and expand Cas9-based genomic imaging.
Tanenbaum et al. (185) combined dCas9-mediated genomic imaging with the SunTag peptide
array to amplify the fluorescent signal generated by each dCas9. Additionally, Ma et al. (72)
developed multicolor genomic imaging using orthogonal dCas9s tagged with different FPs. In
this work, the Sp dCas9, Nm dCas9, and the St1 dCas9 were individually tagged with differently
colored FPs. The different dCas9-FP fusions were targeted to distinct genomic loci by their
corresponding sgRNAs. The authors demonstrated simultaneous tracking dynamics of multiple,
repetitive genomic loci in living cells and that two orthogonal dCas9s may distinguish two genomic

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 245


BI85CH10-Qi ARI 9 May 2016 9:29

loci separated by around 2 mega bases (72). Because the three dCas9s recognize distinct sgRNAs
and PAM sequences, this system provides a tool for tracking the dynamics of multiple genomic
loci simultaneously.
In a method termed CASFISH, fixed cells and tissues can also be efficiently labeled by fluo-
rescently labeled dCas9–sgRNA complexes assembled in vitro (226). Compared with traditional
FISH, which uses high temperature and formamide for denaturation prior to probe hybridization,
CASFISH allows rapid sequence-specific labeling using the dCas9-mediated enzymatic reactions
and, thus, may help to preserve cellular and genomic structure. The authors were able to image
repetitive sequences with a single sgRNA and to image nonrepetitive sequences with an array
of sgRNAs. Given that preassembled Cas9–sgRNA complexes are stable, the authors were also
able to demonstrate multicolor imaging using separately preassembled Cas9–sgRNA complexes
targeting different loci with different colors.
Visualizing the dynamics of fluorescently labeled dCas9 also provides insights into the mecha-
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

nism of how Cas9 searches the genome for its target sites in living cells. Fusing dCas9 to a HaloTag
system and using a single-particle tracking method, Knight et al. (92) found that dCas9 mainly
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

surveys the genome through three-dimensional diffusion, with transient off-target binding. They
also found that dCas9 on-target binding lasts much longer than off-target binding and observed
reduced searching efficiency in heterochromatic regions.

Challenges in Live Cell Genomic Imaging


These programmable sequence-specific DNA-binding proteins are promising tools for moni-
toring genomic dynamics in living cells. However, the visualization of a specific genomic locus
requires recruiting many copies of labeled proteins within a nearby region. How site-specific bind-
ing affects the local chromatin structure and transcriptional activity requires more investigation.
To minimize the perturbation of local chromatin structure, future efforts should focus on reducing
the number of Cas9–sgRNA complexes recruited to a given locus.
Moreover, these techniques require further optimization to efficiently track the dynamics of
any nonrepetitive sequences in living cells. It is challenging to visualize nonrepetitive sequences
using ZF and TALE systems because they require delivery, expression, and recruitment of multiple
DNA-binding proteins, each of which must be directed to a unique sequence. Cas9-based genomic
imaging is more likely to overcome this challenge because it is easier to deliver and express multiple
small-sized sgRNAs. However, the current dCas9 system requires at least 26–36 sgRNAs to
efficiently visualize a nonrepetitive genomic locus (47), which is not trivial in terms of design,
cloning, and delivery and, thus, limits the applications of dCas9 imaging in diverse cell types and
tissues. To improve this system, it is necessary to reduce the required number of sgRNAs and
achieve more sensitive genomic labeling. In addition, more studies are required to evaluate the
specificity and stability of the dCas9-mediated labeling of genomic loci.

Cas9 FOR STUDYING ENDOGENOUS PROTEIN–GENOME


INTERACTIONS AT SPECIFIC LOCI
An affinity-tagged dCas9 can be used in chromatin immunoprecipitation (ChIP) assays to study
proteins that interact with specific portions of the genome (48, 49). The tagged dCas9 can be
targeted to a specific locus and used to pull down the proteins associated with that region. Associ-
ated proteins can then be identified via mass spectrometry or other methods. This CRISPR-based
engineered DNA-binding molecule-mediated ChIP (enChIP) method allows for the study of en-
dogenous protein–genome interactions at specific genomic loci. The authors have successfully

246 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

used this method to characterize proteins interacting with an interferon-γ-responsive promoter


(48, 49).
This CRISPR enChIP method may provide an inexpensive and convenient way of charac-
terizing regional DNA–protein interactions. However, it is necessary to carefully set up proper
controls and compare its results with complementary experiments for validation, given that many
studies have reported extensive genome-wide off-target dCas9-binding events (89, 90, 189, 227,
228). Indeed, ChIP sequencing and other assays have shown that the Cas9–sgRNA complex may
bind from tens to thousands of off-target sites in addition to its target, although most of the com-
plexes bind to the off-target sites with much weaker affinity (89, 90, 189, 227, 228). Of note, only
a small subset of these off-target sites is actually cleaved by Cas9. Detailed analyses have revealed
that the probability of off-target binding events largely depends on chromatin accessibility (89,
90). In this regard, the specificity and efficacy of enChIP pull-down assays may greatly depend
on the choice of sgRNAs and the chromatin accessibility of targeted regions and, therefore, it is
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

necessary to use additional assays to validate the results from CRISPR enChIP.
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

SPATIOTEMPORAL REGULATION OF Cas9 IN GENOMIC EDITING,


REGULATION, AND TARGETING
Tight regulation of Cas9 expression and activity may potentially reduce its off-target effects and,
thus, this is a prerequisite for the future use of Cas9 in clinical applications. Precise spatiotemporal
control of its expression and activity can be achieved at the transcriptional and posttranslational
levels by combining Cas9 with chemical or light-inducible systems. At the transcriptional level,
the expression of Cas9 and nCas9 can be controlled under doxycycline-inducible promoters to
achieve inducible genome editing in cells and mice, although the promoter leakiness may be a
concern in many cell types (138, 229, 230). At the posttranslational level, Davis et al. (158) created
intein-Cas9 fusion proteins that produce active Cas9s when a cell-permeable ligand is present. In
this system, a chemical-responsive intein is inserted into Cas9 to interfere with its nuclease activity.
Upon chemical addition, active Cas9 endonuclease can be produced after intein cleavage through
inducible protein splicing, thereby achieving inducible genome editing at the posttranslational
level. This system enhances the specificity of Cas9-mediated genome editing. However, it should
be noted that in both systems, once active Cas9s are produced, they cannot be removed until
intrinsic protein degradation occurs.
The two components (dCas9 and effector) of the CRISPRi/a gene regulation systems can
also be coupled with chemical and light-inducible systems for spatial and temporal regulation.
For example, Nihongaki et al. (231) and Polstein & Gersbach (232) developed light-inducible
CRISPRa systems to activate endogenous gene expression when cells are exposed to blue light,
thus allowing rapid and reversible control of gene expression. The authors fused light-inducible
heterodimerizing proteins CIB1 and CRY2 to, respectively, dCas9 and a transcriptional activator
(e.g., VP64). Blue-light-induced CIB1–CRY2 dimerization recruits the transcriptional activator
to dCas9 at its target locus, activating gene expression.

Split Cas9s
In addition to light-inducible two-component systems, different versions of split Cas9 have been
created to allow inducible control of gene editing and regulation. For example, Zetsche et al. (233)
split Cas9 into N-terminal and C-terminal fragments and used a rapamycin-binding dimerization
system to induce fragment assembly into active Cas9. Due to their autoassembly, the two split Cas9
fragments need to be spatially separated by tagging with, respectively, nuclear export sequences and

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 247


BI85CH10-Qi ARI 9 May 2016 9:29

nuclear localization sequences. These systems then make use of rapamycin-induced dimerization
to shift the whole complex to the nucleus for DNA targeting. In addition to genomic editing,
inducible gene activation is also achieved with this system by fusing split dCas9 to VP64. A similar
approach has been used to construct split Sa Cas9 for inducible genomic editing (70). Another
strategy reported by Wright et al. (234) uses a split Cas9 that is separated into a nuclease lobe
peptide and an α-helical lobe peptide. An sgRNA is sufficient to induce their assembly into a
whole Cas9 nuclease, although the efficiency of the split Cas9 for genome editing is much lower
than for wild-type Cas9.
Nihongaki et al. (235) created a photoactivatable Cas9 system by fusing split Cas9 fragments
(N713 and C714) with light-inducible dimerization domains. Blue-light-induced dimerization
allowed split Cas9 fragments to reconstitute nuclease activity. This system also allows for optoge-
netic control of nCas9 activity as well as dCas9-mediated transcriptional repression. By tuning the
region and timing of blue light excitation, spatiotemporal and reversible control of Cas9 nuclease
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

activity can be achieved.


Truong et al. (236) developed an intein-mediated split Cas9 system to potentially facilitate in
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

vivo delivery of Cas9. Cas9 was split into N-terminal and C-terminal fragments that were fused
to each of the two components of intein, respectively. The two fragments of Cas9-intein fusions
can be delivered into cells in separate vectors. After intein self-splicing, a full-length Cas9 forms,
exhibiting comparable gene targeting efficiency to wild-type Cas9 in cells. This method allows
the large Sp Cas9 to easily fit into the packaging limit of the recombinant adeno-associated virus
(AAV) system, although its in vivo effectiveness still needs to be assessed.
The development of split Cas9 systems provides new strategies for regulating Cas9 activity,
and it may facilitate the delivery of Cas9 by bypassing the size limitations of some delivery systems.
However, with the exception of intein-mediated split Cas9 (236), most split Cas9 systems have
shown reduced nuclease activity compared with full-length Cas9 (233–235). Additionally, the
background activity of split Cas9 due to autoassociation needs to be critically evaluated.

Cas9 FOR TARGETING RNA


Although most current applications of Cas9 make use of its sequence-specific DNA editing and
targeting capabilities, some articles have opened up the exciting possibility of using Cas9 to target
RNA sequences. Cas9 from the pathogenic Francisella novicida (Fn) can target and degrade mRNA
transcripts for a bacterial lipoprotein, leading to suppression of its host’s immune response (237,
238). For such mRNA recognition, Fn Cas9 forms a complex with its tracrRNA and a novel,
small, CRISPR/Cas-associated RNA (termed a scaRNA) instead of the crRNA. Based on these
observations, Price et al. (50) further engineered this Fn Cas9 system to retarget RNA from
the hepatitis C virus (HCV) in eukaryotic cells. The authors showed that introducing Fn Cas9
together with a synthetically designed, RNA-targeting guide RNA (rgRNA) into cells inhibited
HCV activity, and they suggested that the Fn Cas9–rgRNA complex directly bound to HCV RNA
to inhibit its translation and replication. Interestingly, such Fn Cas9-mediated RNA inhibition
seems to be independent of its nuclease activity and does not require a PAM sequence. Although
the detailed molecular mechanisms of how Fn Cas9–rgRNA binds to and suppresses RNA activity
remain unclear, these findings raise hope that the Cas9 system can be used to target endogenous
RNAs.
Purified Sp Cas9–sgRNA complexes can also target and cleave ssRNA in vitro (51), but this
depends on the presence of a DNA-based PAM-presenting oligonucleotide (PAMmer) that hy-
bridizes with the targeted ssRNA. This system is distinct from the Fn Cas9–rgRNA system, which
does not need a PAM sequence to target RNA.

248 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

The exciting potential of using Cas9 to target RNA may inspire further development of Cas9-
based tools for RNA manipulation. For example, by coupling RNA-targeting Cas9 to different
modulators, it may be possible to regulate the stability, localization, and splicing of the targeted
RNA and to track the real-time dynamics of RNA processing (239).

USE AND DELIVERY OF Cas9


Designing Single Guide RNAs
Designing highly active sgRNAs depends on conforming to a certain set of design rules. To
further improve sgRNA design for nuclease Cas9, Doench et al. (142) made a library of all possible
sgRNAs, tiling across a handful of genes. The authors were able to greatly improve the design of
CRISPR knockout sgRNAs. For example, a typical CRISPR knockout library design before this
study would create 0–2 highly active sgRNAs out of a pool of 6 for 90% of genes; this was improved
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

to 3 or more out of 6 sgRNAs for 90% of genes. Gilbert et al. (40) used a similar technique to
define the rules for effective CRISPRi/a sgRNA, tiling thousands of sgRNAs in a 10 kb window
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

around 49 genes. They defined a set of rules from which it is possible to obtain at least 1–2 highly
active (80–99% repression) sgRNAs out of a pool of 5–10. Currently, it is usually necessary to
clone a handful of sgRNAs per target gene, as there are other factors affecting sgRNA function
that have not yet been defined.
Although there are several online tools available to aid in designing sgRNAs, it is also possible
to design sgRNAs manually. Several groups have distilled information from large data sets to come
up with a set of simple rules for what makes an sgRNA most effective in various contexts. Design
rules include, for example, the following:
 All sgRNAs must be adjacent to a PAM site: Sp Cas9 uses NGG or a less efficient NAG
(3, 68). Sa Cas9’s PAM is NNGRRT (69).
 When using a U6 promoter for sgRNA expression in mammalian cells, the first nucleotide
of the sgRNA must be a G for effective expression, although it has been suggested that an A
may work in some contexts (240). It has been shown with Cas9 nuclease that an sgRNA can
still function if the 5 G of the sgRNA is mismatched with the target site, which is useful if
no target site can be found where a G is 18–25 bp upstream of NGG (139, 241).
 An sgRNA expressed from a U6 promoter in mammalian cells should not contain a stretch
of four or more uracils (U’s) in a row or it will be terminated prematurely due to the activity
of RNA polymerase III (242). A stretch of U’s near the 3 end of the guide sequence is
unfavorable for Cas9–sgRNA binding (138).
 There are mixed reports about the effect of GC content. An article by Wang et al. (138) sug-
gested that sgRNAs with a very high or low GC content were less effective when combined
with nuclease Cas9. Another study, by Gilbert et al. (40), used dCas9 fused to effectors and
found that variations in GC content did not significantly change sgRNA effectiveness.
 Long stretches of the same nucleotide greatly decrease sgRNA activity (40).

Although great strides have been made in finding ever better sgRNAs, the design rules for both
Cas9-based genome editing and gene regulation will no doubt benefit from further optimization
in the future.

Choosing Target Sites


When using nuclease Cas9 to knock out a gene through the creation of indels, it is most common
to target an early exon in the coding sequence, to disrupt as much of the protein as possible by

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 249


BI85CH10-Qi ARI 9 May 2016 9:29

a frameshift mutation (138, 139). To determine the most potent targeting sites for CRISPRi/a,
Gilbert et al. (40) and Konermann et al. (180) have tiled sgRNAs around the promoters of various
genes. By tiling a library of sgRNAs in a 10 kb window around the TSS of 49 genes (approximately
55,000 sgRNAs total), Gilbert et al. (40) systematically examined which sites allowed for the most
active sgRNAs for both CRISPRa and CRISPRi. CRISPRi is most effective with sgRNAs in
the −50 to +300 bp window around the TSS, with the absolute highest repression occurring in
approximately the 50–100 bp window downstream of the TSS. The most effective sgRNAs for
CRISPRa are targeted to a window −400 to −50 bp upstream of the TSS. Both of these activity
windows are consistent with data regarding the mechanism of action for the KRAB repressor (243,
244) and VP64 activator (245).
Although these studies have been greatly informative for the creation of effective sgRNAs,
much work still remains for figuring out the full set of rules, particularly for different cell types
and organisms. Future work should systematically determine how local chromatin, transcriptional,
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

and epigenetic status (i.e., local histone marks, methylated DNA, or actively transcribing DNA)
affect genome targeting, binding, nuclease, and regulatory activities. For example, there is a positive
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

correlation between the level of sgRNA expression in mammalian cells and the regulatory function
of Cas9, suggesting that the sgRNA expression level is likely a limiting factor for the dCas9 function
(7). For CRISPRi/a, the absolute dosage of sgRNAs and dCas9, as well as the most effective ratio
of the two, also require future study.

Delivery Methods
There are many vectors available on Addgene (https://www.addgene.org/) and through com-
mercial vendors that encode various CRISPR components. The particular components and their
delivery will depend on the assay and cell type. After picking an appropriate Cas9-encoding vector,
a complementary sgRNA-encoding vector can also be adopted. Directing a dCas9 effector to a
sequence of interest requires the simple cloning of a short sgRNA (approximately 100 bp) and
then introducing both components into the cell type of interest (141, 174, 188). For CRISPRi/a, it
is standard to also use one or multiple nontargeting sgRNAs—which should not target anywhere
in the genome or introduced DNA—as a negative control and as a way to normalize results.
Depending on the desired application, there are various ways in which the Cas9 system can
be introduced. For an application such as a pooled screen, it is often desirable to make a stable
Cas9 or dCas9 effector cell line using lentiviral or retroviral vectors before the introduction of a
lentiviral sgRNA pool (138). This method may be problematic when working with primary cells.
Alternatively, it is possible to use vectors that encode both Cas9 and a single sgRNA (139, 141).
It is necessary to have just one sgRNA present per cell in pooled screens because this gives the
highest signal-to-noise ratio when correlating an observed phenotype with the effect of a given
sgRNA. Thus, viral delivery is ideal for pooled screens because the viral particles can be titered
such that, on average, there is less than one sgRNA-containing viral particle per cell in the infected
pool (giving a multiplicity of infection of less than 1).
For shorter-term or smaller-scale experiments in cell lines, it is also possible to use transient
transfection to introduce plasmids containing Cas9 or dCas9 effectors and sgRNAs. When knock-
ing out or editing a gene or handful of genes, this may be ideal, as continued expression of the
CRISPR components is not necessary once the desired genome modification has occurred. For
CRISPRi/a, transfection is best suited to activation assays, which can often be assessed after 24–
48 hours (37). Repression assays may be less effective when using transfection because CRISPRi
affects transcriptional efficiency and not mRNA or protein stability. The already-present mRNA
and protein for the target gene must be degraded at its normal turnover rate, which can vary from

250 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

gene to gene, before the full effects of repression are seen. It may take several days to see repression,
but by that time the CRISPRi plasmids may have been diluted out of the cells as they divide and,
thus, CRISPRi may become less effective. Thus, any studies using CRISPRi may be better suited
to viral delivery to maintain CRISPRi component expression over a longer time scale. With an
eye toward the technical aspects of Cas9–sgRNA component expression and delivery, it is quite
straightforward to design an effective assay.
For gene therapy, the delivery of the Cas9 system is more challenging and raises more safety
concerns. AAV-based vectors are the preferred candidates for somatic gene therapy due to their
mild immune response, lack of pathogenicity, and ability to target nondividing cells (246). How-
ever, the coding sequences of Sp Cas9, the most widely used Cas9, and its sgRNA are already
approaching the packaging limit of AAV-based gene therapy vectors. The large size of Sp Cas9
makes its usage in AAV-based gene therapy unfavorable, especially when promoter sequences,
localization signals, donor DNA, or additional sgRNAs are needed. The Sa Cas9 has an ap-
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

proximately 1 kb shorter coding sequence than Sp Cas9, allowing it to be easily packaged into
AAV-based vectors for gene editing (69, 247). Using this AAV Sa Cas9 approach, Ran et al. (69)
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

efficiently induced indel formation of more than 40% of the cholesterol regulatory Pcsk9 (propro-
tein convertase subtilisin/kexin type 9) gene in mouse liver after 1 week of treatment, causing a
95% drop in serum Pcsk9 and a 40% drop in total cholesterol.
Besides viral systems, nonviral methods have been used to deliver Cas9-encoding plasmids or
Cas9 mRNA into animal cells and tissues. For example, coinjection of Cas9-encoding mRNA and
an sgRNA into zygotes or embryos can generate genome-edited animals (36, 128, 248, 249). Cas9-
containing vectors can also be introduced into adult animals through hydrodynamic injections,
leading to efficient gene corrections or mutations (125, 250).
The use of purified Cas9–sgRNA ribonucleoproteins (RNPs) is another option for cellular
delivery. Compared with a viral or nonviral nucleotide delivery method, the RNP delivery systems
allows for fast action of the RNP complex in the nucleus and a shorter duration of the Cas9
nuclease presence in the cells and, thus, it may increase the efficiency and reduce the off-target
effects. RNP delivery also avoids undesired genomic alterations that can occur when using other
nucleotide delivery methods.
RNP delivery of Cas9 and sgRNAs can be achieved through a variety of strategies. Many
methods traditionally used for nucleotide transfection have been proven useful for Cas9 RNP
delivery, including microinjections, electroporation, and lipid-mediated transfection. Microinjec-
tion of purified Cas9 RNP complexes into animal embryos successfully generated genome-edited
animals (251, 252). Electroporation methods have been established to introduce Cas9–sgRNA
complexes into primary cells and embryonic stem cells, and these can induce targeted gene muta-
tions and large chromosome deletions while minimizing the off-target effects (108, 161, 253–256).
Commercially available nucleotide transfection reagents have also been adapted to deliver Cas9–
sgRNA RNP complexes (256, 257). Given that the Cas9–sgRNA complex is inherently anionic,
Zuris et al. (257) adapted cationic lipid transfection reagents to deliver the Cas9 nuclease, nCas9,
and the dCas9-VP64 transcriptional activator. This method allowed for gene modification of up
to 80% in cultured cells and approximately 20% in mouse inner ear hair cells in vivo (257).
Other approaches have also been explored for delivering RNP complexes. For example,
Ramakrishna et al. (258) used cell-penetrating peptides (CPPs) to induce gene editing through
Cas9 conjugated with CPPs and through the sgRNA in complex with CPPs. These CPP com-
plexes entered cells in the form of positively charged nanoparticles to achieve gene editing in a
variety of human cell types. D’Astolfo et al. (259) developed a method of induced transduction by
osmocytosis and propanebetaine (termed iTOP) to deliver Cas9–sgRNA complexes, leading to
highly efficient gene editing in primary cells. This method allows uptake and release of proteins

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 251


BI85CH10-Qi ARI 9 May 2016 9:29

Figure 3 TOOLS AND APPLICATIONS BASED ON Cas9 AND nCas9


Applications of
Cas9 nuclease or nickase
CRISPR-associated a Genome editing
protein 9 (Cas9) in
gene editing, gene
regulation, epigenome
editing, and genomic
imaging. (a) Cas9- Targeted gene mutagenesis/sequence replacement
mediated site-specific
Large-scale chromosomal rearrangment
genome editing has a
Genome-scale gene knockout screening
variety of applications.
Generation of transgenic organisms
(b–d ) Tools based on
nuclease deactivated Disease modeling
Cas9 (dCas9) are built Gene therapy
to achieve (b) site-
specific gene
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

TOOLS AND APPLICATIONS BASED ON dCas9


repression or
activation [i.e., b Gene repression/activation Repressor
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

CRISPR interference (CRISPRi/a) Activator


(CRISPRi) and dCas9
CRISPR activation
(CRISPRa)],
(c) epigenetic Target gene
modification, or
(d ) genomic imaging. sgRNA Targeted gene regulation
Their applications are Repressor Genome-scale gene activation/
repression screening
summarized. RNA aptamer Activator Orthogonal gene regulation
Transcriptional
RBP
activators and
repressors, epigenetic
modifiers, and Target gene
fluorescent proteins
scRNA
can be fused or
directed to either (b–d )
the dCas9 or (b) the c Epigenome editing Epigenetic modifier Modification of
epigenetic marks near
single guide RNA dCas9 targeted site
(sgRNA) bound to Regulation of downstream
their target DNA site. gene when targeted to
Abbreviations: RBP, enhancer/promoter
RNA-binding Target gene
proteins; scRNA,
scaffold RNA.

d Genomic imaging
Fluorescent protein Imaging of repetitive/
nonrepetitive genomic
loci targeted by dCas9
Live/fixed cell imaging

252 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

and other molecules in a variety of primary cells through the macropinocytosis pathway. In addi-
tion, Sun et al. (260) made use of a DNA nanoparticle, termed a nanoclew, to deliver a Cas9 RNP
complex into U2OS cells for gene editing. They also used nanoclews to deliver the Cas9–sgRNA
complex in vivo, successfully editing a genome-integrated GFP reporter in 25% of U2OS tumor
cells, which had been xenografted in mice.
Compared with vector-mediated nucleotide delivery methods, Cas9-mediated genome editing
via RNP delivery methods (108, 161, 253, 257, 258) exhibits higher fidelity and lower cell toxicity,
bypasses the safety problems of introducing foreign DNA into the host genome and, thus, may
provide a platform for developing gene therapy tools.

CONCLUSIONS AND FUTURE PERSPECTIVES


Although we have not yet harnessed the full potential of CRISPR/Cas9, this technology has
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

brought forth revolutionary changes in genomic research, including genome editing, regulation,
and imaging (Figure 3). As we look to the future, we can envision the advances that CRISPR/Cas9
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

technology will bring to basic bioscience researchers and clinicians. Basic scientists are already
making great strides in understanding and manipulating biology using CRISPR/Cas9 technology.
In the clinic, we can look forward to new therapies for genetic diseases (using Cas9 genome editing
or CRISPRi/a) and new diagnostic techniques (using dCas9-based imaging).
Basic research similar to that which uncovered CRISPR/Cas9 allows us to make use of nature’s
technological toolbox, which has been honed for billions of years through evolution. Although
most research thus far has focused on the type II Cas9 proteins, there is much still to be discovered
about the broader CRISPR systems from other diverse species of bacteria and Archaea. New
CRISPR systems, hidden in plain sight in the genomes of the organisms around us, may continue
to surprise us in the future with their elegant mechanisms and function, offering powerful and
groundbreaking technologies.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
The authors thank all the members of the Qi lab, as well as Dr. Fuguo Jiang and Arjun Aditham
for advice and helpful discussions. Thanks also to Matias Kaplan for preparing the structure
comparison of Cas9s. L.S.Q. acknowledges support from National Institutes for Health’s Office
of the Director and the National Institute of Dental and Craniofacial Research. M.L. acknowledges
support from an Agilent University Relations grant. This work was supported by NIH grants R01
DA036858 and U01 EB021240 (H.W., L.S.Q.) and DP5 OD017887 (M.L., L.S.Q.).

LITERATURE CITED
1. Ishino Y, Shinagawa H, Makino K, Amemura M, Nakata A. 1987. Nucleotide sequence of the iap gene,
responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene
product. J. Bacteriol. 169:5429–33
2. Jansen R, Embden JD, Gaastra W, Schouls LM. 2002. Identification of genes that are associated with
DNA repeats in prokaryotes. Mol. Microbiol. 43:1565–75

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 253


BI85CH10-Qi ARI 9 May 2016 9:29

3. Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. 2012. A programmable dual-
RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337:816–21
4. Gasiunas G, Barrangou R, Horvath P, Siksnys V. 2012. Cas9-crRNA ribonucleoprotein complex medi-
ates specific DNA cleavage for adaptive immunity in bacteria. PNAS 109:E2579–86
5. Cong L, Ran FA, Cox D, Lin S, Barretto R, et al. 2013. Multiplex genome engineering using CRISPR/Cas
systems. Science 339:819–23
6. Mali P, Yang L, Esvelt KM, Aach J, Guell M, et al. 2013. RNA-guided human genome engineering via
Cas9. Science 339:823–26
7. Jinek M, East A, Cheng A, Lin S, Ma E, Doudna J. 2013. RNA-programmed genome editing in human
cells. eLife 2:e00471
8. Mali P, Esvelt KM, Church GM. 2013. Cas9 as a versatile tool for engineering biology. Nat. Methods
10:957–63
9. Hsu PD, Lander ES, Zhang F. 2014. Development and applications of CRISPR-Cas9 for genome
engineering. Cell 157:1262–78
10. Doudna JA, Charpentier E. 2014. The new frontier of genome engineering with CRISPR-Cas9. Science
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

346:1258096
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

11. Mojica FJ, Diez-Villaseñor C, Soria E, Juez G. 2000. Biological significance of a family of regularly
spaced repeats in the genomes of Archaea, Bacteria and mitochondria. Mol. Microbiol. 36:244–46
12. Barrangou R, Fremaux C, Deveau H, Richards M, Boyaval P, et al. 2007. CRISPR provides acquired
resistance against viruses in prokaryotes. Science 315:1709–12
13. Garneau JE, Dupuis ME, Villion M, Romero DA, Barrangou R, et al. 2010. The CRISPR/Cas bacterial
immune system cleaves bacteriophage and plasmid DNA. Nature 468:67–71
14. Deltcheva E, Chylinski K, Sharma CM, Gonzales K, Chao Y, et al. 2011. CRISPR RNA maturation by
trans-encoded small RNA and host factor RNase III. Nature 471:602–7
15. Rouet P, Smih F, Jasin M. 1994. Introduction of double-strand breaks into the genome of mouse cells
by expression of a rare-cutting endonuclease. Mol. Cell. Biol. 14:8096–106
16. Rouet P, Smih F, Jasin M. 1994. Expression of a site-specific endonuclease stimulates homologous
recombination in mammalian cells. PNAS 91:6064–68
17. Rudin N, Sugarman E, Haber JE. 1989. Genetic and physical analysis of double-strand break repair and
recombination in Saccharomyces cerevisiae. Genetics 122:519–34
18. Plessis A, Perrin A, Haber JE, Dujon B. 1992. Site-specific recombination determined by I-SceI, a
mitochondrial group I intron-encoded endonuclease expressed in the yeast nucleus. Genetics 130:451–60
19. Choulika A, Perrin A, Dujon B, Nicolas JF. 1995. Induction of homologous recombination in mammalian
chromosomes by using the I-SceI system of Saccharomyces cerevisiae. Mol. Cell. Biol. 15:1968–73
20. Porteus M. 2016. Genome editing: a new approach to human therapeutics. Annu. Rev. Pharmacol. Toxicol.
56:163–90
21. Silva G, Poirot L, Galetto R, Smith J, Montoya G, et al. 2011. Meganucleases and other tools for targeted
genome engineering: perspectives and challenges for gene therapy. Curr. Gene Ther. 11:11–27
22. Grizot S, Epinat JC, Thomas S, Duclert A, Rolland S, et al. 2010. Generation of redesigned homing
endonucleases comprising DNA-binding domains derived from two different scaffolds. Nucleic Acids Res.
38:2006–18
23. Miller J, McLachlan AD, Klug A. 1985. Repetitive zinc-binding domains in the protein transcription
factor IIIA from Xenopus oocytes. EMBO J. 4:1609–14
24. Kim YG, Cha J, Chandrasegaran S. 1996. Hybrid restriction enzymes: zinc finger fusions to Fok I
cleavage domain. PNAS 93:1156–60
25. Urnov FD, Rebar EJ, Holmes MC, Zhang HS, Gregory PD. 2010. Genome editing with engineered
zinc finger nucleases. Nat. Rev. Genet. 11:636–46
26. Boch J, Scholze H, Schornack S, Landgraf A, Hahn S, et al. 2009. Breaking the code of DNA binding
specificity of TAL-type III effectors. Science 326:1509–12
27. Christian M, Cermak T, Doyle EL, Schmidt C, Zhang F, et al. 2010. Targeting DNA double-strand
breaks with TAL effector nucleases. Genetics 186:757–61
28. Joung JK, Sander JD. 2013. TALENs: a widely applicable technology for targeted genome editing.
Nat. Rev. Mol. Cell Biol. 14:49–55

254 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

29. Marraffini LA, Sontheimer EJ. 2008. CRISPR interference limits horizontal gene transfer in staphylo-
cocci by targeting DNA. Science 322:1843–45
30. Bolotin A, Quinquis B, Sorokin A, Ehrlich SD. 2005. Clustered regularly interspaced short palindrome
repeats (CRISPRs) have spacers of extrachromosomal origin. Microbiology 151:2551–61
31. Jiang W, Bikard D, Cox D, Zhang F, Marraffini LA. 2013. RNA-guided editing of bacterial genomes
using CRISPR-Cas systems. Nat. Biotechnol. 31:233–39
32. DiCarlo JE, Norville JE, Mali P, Rios X, Aach J, Church GM. 2013. Genome engineering in Saccha-
romyces cerevisiae using CRISPR-Cas systems. Nucleic Acids Res. 41:4336–43
33. Li JF, Norville JE, Aach J, McCormack M, Zhang D, et al. 2013. Multiplex and homologous
recombination-mediated genome editing in Arabidopsis and Nicotiana benthamiana using guide RNA
and Cas9. Nat. Biotechnol. 31:688–91
34. Nekrasov V, Staskawicz B, Weigel D, Jones JD, Kamoun S. 2013. Targeted mutagenesis in the model
plant Nicotiana benthamiana using Cas9 RNA-guided endonuclease. Nat. Biotechnol. 31:691–93
35. Sander JD, Joung JK. 2014. CRISPR-Cas systems for editing, regulating and targeting genomes.
Nat. Biotechnol. 32:347–55
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

36. Wang H, Yang H, Shivalila CS, Dawlaty MM, Cheng AW, et al. 2013. One-step generation of mice
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell 153:910–18


37. Gilbert LA, Larson MH, Morsut L, Liu Z, Brar GA, et al. 2013. CRISPR-mediated modular RNA-
guided regulation of transcription in eukaryotes. Cell 154:442–51
38. Hilton IB, D’Ippolito AM, Vockley CM, Thakore PI, Crawford GE, et al. 2015. Epigenome editing by
a CRISPR-Cas9-based acetyltransferase activates genes from promoters and enhancers. Nat. Biotechnol.
33:510–17
39. Kearns NA, Pham H, Tabak B, Genga RM, Silverstein NJ, et al. 2015. Functional annotation of native
enhancers with a Cas9-histone demethylase fusion. Nat. Methods 12:401–3
40. Gilbert LA, Horlbeck MA, Adamson B, Villalta JE, Chen Y, et al. 2014. Genome-scale CRISPR-mediated
control of gene repression and activation. Cell 159:647–61
41. Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, et al. 2013. Repurposing CRISPR as an
RNA-guided platform for sequence-specific control of gene expression. Cell 152:1173–83
42. Lindhout BI, Fransz P, Tessadori F, Meckel T, Hooykaas PJ, van der Zaal BJ. 2007. Live cell imaging
of repetitive DNA sequences via GFP-tagged polydactyl zinc finger proteins. Nucleic Acids Res. 35:e107
43. Miyanari Y, Ziegler-Birling C, Torres-Padilla ME. 2013. Live visualization of chromatin dynamics with
fluorescent TALEs. Nat. Struct. Mol. Biol. 20:1321–24
44. Ma H, Reyes-Gutierrez P, Pederson T. 2013. Visualization of repetitive DNA sequences in human
chromosomes with transcription activator-like effectors. PNAS 110:21048–53
45. Thanisch K, Schneider K, Morbitzer R, Solovei I, Lahaye T, et al. 2014. Targeting and tracing of specific
DNA sequences with dTALEs in living cells. Nucleic Acids Res. 42:e38
46. Pederson T. 2014. Repeated TALEs: visualizing DNA sequence localization and chromosome dynamics
in live cells. Nucleus 5:28–31
47. Chen B, Gilbert LA, Cimini BA, Schnitzbauer J, Zhang W, et al. 2013. Dynamic imaging of genomic
loci in living human cells by an optimized CRISPR/Cas system. Cell 155:1479–91
48. Fujita T, Fujii H. 2014. Identification of proteins associated with an IFNγ-responsive promoter by a
retroviral expression system for enChIP using CRISPR. PLOS ONE 9:e103084
49. Fujita T, Fujii H. 2013. Efficient isolation of specific genomic regions and identification of associated
proteins by engineered DNA-binding molecule-mediated chromatin immunoprecipitation (enChIP)
using CRISPR. Biochem. Biophys. Res. Commun. 439:132–36
50. Price AA, Sampson TR, Ratner HK, Grakoui A, Weiss DS. 2015. Cas9-mediated targeting of viral RNA
in eukaryotic cells. PNAS 112:6164–69
51. O’Connell MR, Oakes BL, Sternberg SH, East-Seletsky A, Kaplan M, Doudna JA. 2014. Programmable
RNA recognition and cleavage by CRISPR/Cas9. Nature 516:263–66
52. Mojica FJ, Diez-Villasenor C, Garcia-Martinez J, Soria E. 2005. Intervening sequences of regularly
spaced prokaryotic repeats derive from foreign genetic elements. J. Mol. Evol. 60:174–82
53. Rath D, Amlinger L, Rath A, Lundgren M. 2015. The CRISPR-Cas immune system: biology, mecha-
nisms and applications. Biochimie 117:119–28

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 255


BI85CH10-Qi ARI 9 May 2016 9:29

54. Koonin EV, Krupovic M. 2015. Evolution of adaptive immunity from transposable elements combined
with innate immune systems. Nat. Rev. Genet. 16:184–92
55. Pourcel C, Salvignol G, Vergnaud G. 2005. CRISPR elements in Yersinia pestis acquire new repeats
by preferential uptake of bacteriophage DNA, and provide additional tools for evolutionary studies.
Microbiology 151:653–63
56. Brouns SJ, Jore MM, Lundgren M, Westra ER, Slijkhuis RJ, et al. 2008. Small CRISPR RNAs guide
antiviral defense in prokaryotes. Science 321:960–64
57. Makarova KS, Wolf YI, Alkhnbashi OS, Costa F, Shah SA, et al. 2015. An updated evolutionary classi-
fication of CRISPR-Cas systems. Nat. Rev. Microbiol. 13:722–36
58. Zetsche B, Gootenberg JS, Abudayyeh OO, Slaymaker IM, Makarova KS, et al. 2015. Cpf1 is a single
RNA-guided endonuclease of a class 2 CRISPR-Cas system. Cell 163:759–71
59. Shmakov S, Abudayyeh OO, Makarova KS, Wolf YI, Gootenberg JS, et al. 2015. Discovery and functional
characterization of diverse class 2 CRISPR-Cas systems. Mol. Cell 60:385–97
60. Wiedenheft B, Sternberg SH, Doudna JA. 2012. RNA-guided genetic silencing systems in bacteria and
archaea. Nature 482:331–38
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

61. Barrangou R, Marraffini LA. 2014. CRISPR-Cas systems: Prokaryotes upgrade to adaptive immunity.
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Mol. Cell 54:234–44


62. Bondy-Denomy J, Davidson AR. 2014. To acquire or resist: the complex biological effects of CRISPR-
Cas systems. Trends Microbiol. 22:218–25
63. Mojica FJ, Diez-Villasenor C, Garcia-Martinez J, Almendros C. 2009. Short motif sequences determine
the targets of the prokaryotic CRISPR defence system. Microbiology 155:733–40
64. Deveau H, Barrangou R, Garneau JE, Labonte J, Fremaux C, et al. 2008. Phage response to CRISPR-
encoded resistance in Streptococcus thermophilus. J. Bacteriol. 190:1390–400
65. Shah SA, Erdmann S, Mojica FJ, Garrett RA. 2013. Protospacer recognition motifs: mixed identities
and functional diversity. RNA Biol. 10:891–99
66. Hale CR, Zhao P, Olson S, Duff MO, Graveley BR, et al. 2009. RNA-guided RNA cleavage by a CRISPR
RNA-Cas protein complex. Cell 139:945–56
67. Fonfara I, Le Rhun A, Chylinski K, Makarova KS, Lecrivain AL, et al. 2014. Phylogeny of Cas9 de-
termines functional exchangeability of dual-RNA and Cas9 among orthologous type II CRISPR-Cas
systems. Nucleic Acids Res. 42:2577–90
68. Hsu PD, Scott DA, Weinstein JA, Ran FA, Konermann S, et al. 2013. DNA targeting specificity of
RNA-guided Cas9 nucleases. Nat. Biotechnol. 31:827–32
69. Ran FA, Cong L, Yan WX, Scott DA, Gootenberg JS, et al. 2015. In vivo genome editing using Staphy-
lococcus aureus Cas9. Nature 520:186–91
70. Nishimasu H, Cong L, Yan WX, Ran FA, Zetsche B, et al. 2015. Crystal structure of Staphylococcus aureus
Cas9. Cell 162:1113–26
71. Esvelt KM, Mali P, Braff JL, Moosburner M, Yaung SJ, Church GM. 2013. Orthogonal Cas9 proteins
for RNA-guided gene regulation and editing. Nat. Methods 10:1116–21
72. Ma H, Naseri A, Reyes-Gutierrez P, Wolfe SA, Zhang S, Pederson T. 2015. Multicolor CRISPR labeling
of chromosomal loci in human cells. PNAS 112:3002–7
73. Horvath P, Romero DA, Coute-Monvoisin AC, Richards M, Deveau H, et al. 2008. Diversity, activity,
and evolution of CRISPR loci in Streptococcus thermophilus. J. Bacteriol. 190:1401–12
74. Zhang Y, Heidrich N, Ampattu BJ, Gunderson CW, Seifert HS, et al. 2013. Processing-independent
CRISPR RNAs limit natural transformation in Neisseria meningitidis. Mol. Cell 50:488–503
75. Briner AE, Donohoue PD, Gomaa AA, Selle K, Slorach EM, et al. 2014. Guide RNA functional modules
direct Cas9 activity and orthogonality. Mol. Cell 56:333–39
76. Mali P, Aach J, Stranges PB, Esvelt KM, Moosburner M, et al. 2013. CAS9 transcriptional activators
for target specificity screening and paired nickases for cooperative genome engineering. Nat. Biotechnol.
31:833–38
77. Ran FA, Hsu PD, Lin CY, Gootenberg JS, Konermann S, et al. 2013. Double nicking by RNA-guided
CRISPR Cas9 for enhanced genome editing specificity. Cell 154:1380–89
78. Shen B, Zhang W, Zhang J, Zhou J, Wang J, et al. 2014. Efficient genome modification by CRISPR-
Cas9 nickase with minimal off-target effects. Nat. Methods 11:399–402

256 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

79. Trevino AE, Zhang F. 2014. Genome editing using Cas9 nickases. Methods Enzymol. 546:161–74
80. Cho SW, Kim S, Kim Y, Kweon J, Kim HS, et al. 2014. Analysis of off-target effects of CRISPR/Cas-
derived RNA-guided endonucleases and nickases. Genome Res. 24:132–41
81. Jinek M, Jiang F, Taylor DW, Sternberg SH, Kaya E, et al. 2014. Structures of Cas9 endonucleases
reveal RNA-mediated conformational activation. Science 343:1247997
82. Jiang F, Zhou K, Ma L, Gressel S, Doudna JA. 2015. A Cas9-guide RNA complex preorganized for
target DNA recognition. Science 348:1477–81
83. Nishimasu H, Ran FA, Hsu PD, Konermann S, Shehata SI, et al. 2014. Crystal structure of Cas9 in
complex with guide RNA and target DNA. Cell 156:935–49
84. Anders C, Niewoehner O, Duerst A, Jinek M. 2014. Structural basis of PAM-dependent target DNA
recognition by the Cas9 endonuclease. Nature 513:569–73
85. Jiang F, Doudna JA. 2015. The structural biology of CRISPR-Cas systems. Curr. Opin. Struct. Biol.
30:100–11
86. Sternberg SH, LaFrance B, Kaplan M, Doudna JA. 2015. Conformational control of DNA target cleavage
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

by CRISPR-Cas9. Nature 527:110–13


87. Sternberg SH, Redding S, Jinek M, Greene EC, Doudna JA. 2014. DNA interrogation by the CRISPR
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

RNA-guided endonuclease Cas9. Nature 507:62–67


88. Szczelkun MD, Tikhomirova MS, Sinkunas T, Gasiunas G, Karvelis T, et al. 2014. Direct observation
of R-loop formation by single RNA-guided Cas9 and Cascade effector complexes. PNAS 111:9798–803
89. O’Geen H, Henry IM, Bhakta MS, Meckler JF, Segal DJ. 2015. A genome-wide analysis of Cas9 binding
specificity using ChIP-seq and targeted sequence capture. Nucleic Acids Res. 43:3389–404
90. Wu X, Scott DA, Kriz AJ, Chiu AC, Hsu PD, et al. 2014. Genome-wide binding of the CRISPR
endonuclease Cas9 in mammalian cells. Nat. Biotechnol. 32:670–76
91. Josephs EA, Kocak DD, Fitzgibbon CJ, McMenemy J, Gersbach CA, Marszalek PE. 2015. Structure
and specificity of the RNA-guided endonuclease Cas9 during DNA interrogation, target binding and
cleavage. Nucleic Acids Res. 43:8924–41
92. Knight SC, Xie L, Deng W, Guglielmi B, Witkowsky LB, et al. 2015. Dynamics of CRISPR-Cas9
genome interrogation in living cells. Science 350:823–26
93. Lieber MR. 2010. The mechanism of double-strand DNA break repair by the nonhomologous DNA
end-joining pathway. Annu. Rev. Biochem. 79:181–211
94. Xiao-Jie L, Hui-Ying X, Zun-Ping K, Jin-Lian C, Li-Juan J. 2015. CRISPR-Cas9: a new and promising
player in gene therapy. J. Med. Genet. 52:289–96
95. Kleinstiver BP, Prew MS, Tsai SQ, Nguyen NT, Topkar VV, et al. 2015. Broadening the targeting
range of Staphylococcus aureus CRISPR-Cas9 by modifying PAM recognition. Nat. Biotechnol. 33:1293–98
96. Kleinstiver BP, Prew MS, Tsai SQ, Topkar VV, Nguyen NT, et al. 2015. Engineered CRISPR-Cas9
nucleases with altered PAM specificities. Nature 523:481–85
97. Xie K, Yang Y. 2013. RNA-guided genome editing in plants using a CRISPR-Cas system. Mol. Plant
6:1975–83
98. Jiang W, Zhou H, Bi H, Fromm M, Yang B, Weeks DP. 2013. Demonstration of CRISPR/Cas9/sgRNA-
mediated targeted gene modification in Arabidopsis, tobacco, sorghum and rice. Nucleic Acids Res. 41:e188
99. Niu Y, Shen B, Cui Y, Chen Y, Wang J, et al. 2014. Generation of gene-modified cynomolgus monkey
via Cas9/RNA-mediated gene targeting in one-cell embryos. Cell 156:836–43
100. Gratz SJ, Cummings AM, Nguyen JN, Hamm DC, Donohue LK, et al. 2013. Genome engineering of
Drosophila with the CRISPR RNA-guided Cas9 nuclease. Genetics 194:1029–35
101. Liu P, Long L, Xiong K, Yu B, Chang N, et al. 2014. Heritable/conditional genome editing in C. elegans
using a CRISPR-Cas9 feeding system. Cell Res. 24:886–89
102. Friedland AE, Tzur YB, Esvelt KM, Colaiacovo MP, Church GM, Calarco JA. 2013. Heritable genome
editing in C. elegans via a CRISPR-Cas9 system. Nat. Methods 10:741–43
103. Dow LE. 2015. Modeling disease in vivo with CRISPR/Cas9. Trends Mol. Med. 21:609–21
104. Blasco RB, Karaca E, Ambrogio C, Cheong TC, Karayol E, et al. 2014. Simple and rapid in vivo
generation of chromosomal rearrangements using CRISPR/Cas9 technology. Cell Rep. 9:1219–27

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 257


BI85CH10-Qi ARI 9 May 2016 9:29

105. Canver MC, Bauer DE, Dass A, Yien YY, Chung J, et al. 2014. Characterization of genomic deletion
efficiency mediated by clustered regularly interspaced palindromic repeats (CRISPR)/Cas9 nuclease
system in mammalian cells. J. Biol. Chem. 289:21312–24
106. He Z, Proudfoot C, Mileham AJ, McLaren DG, Whitelaw CB, Lillico SG. 2015. Highly efficient targeted
chromosome deletions using CRISPR/Cas9. Biotechnol. Bioeng. 112:1060–64
107. Choi PS, Meyerson M. 2014. Targeted genomic rearrangements using CRISPR/Cas technology.
Nat. Commun. 5:3728
108. Kim S, Kim D, Cho SW, Kim J, Kim JS. 2014. Highly efficient RNA-guided genome editing in human
cells via delivery of purified Cas9 ribonucleoproteins. Genome Res. 24:1012–19
109. Kraft K, Geuer S, Will AJ, Chan WL, Paliou C, et al. 2015. Deletions, inversions, duplications: engi-
neering of structural variants using CRISPR/Cas in mice. Cell Rep. pii:S2211–47
110. Maddalo D, Manchado E, Concepcion CP, Bonetti C, Vidigal JA, et al. 2014. In vivo engineering of
oncogenic chromosomal rearrangements with the CRISPR/Cas9 system. Nature 516:423–27
111. Torres R, Martin MC, Garcia A, Cigudosa JC, Ramirez JC, Rodriguez-Perales S. 2014. Engineering
human tumour-associated chromosomal translocations with the RNA-guided CRISPR-Cas9 system.
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Nat. Commun. 5:3964


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

112. Barrangou R, May AP. 2015. Unraveling the potential of CRISPR-Cas9 for gene therapy. Expert Opin.
Biol. Ther. 15:311–14
113. Ebina H, Misawa N, Kanemura Y, Koyanagi Y. 2013. Harnessing the CRISPR/Cas9 system to disrupt
latent HIV-1 provirus. Sci. Rep. 3:2510
114. Hu W, Kaminski R, Yang F, Zhang Y, Cosentino L, et al. 2014. RNA-directed gene editing specifically
eradicates latent and prevents new HIV-1 infection. PNAS 111:11461–66
115. Liu X, Hao R, Chen S, Guo D, Chen Y. 2015. Inhibition of hepatitis B virus by CRISPR/Cas9 system
via targeting the conserved regions of viral genome. J. Gen. Virol. 96:2252–61
116. Dong C, Qu L, Wang H, Wei L, Dong Y, Xiong S. 2015. Targeting hepatitis B virus cccDNA by
CRISPR/Cas9 nuclease efficiently inhibits viral replication. Antivir. Res. 118:110–17
117. Seeger C, Sohn JA. 2014. Targeting hepatitis B virus with CRISPR/Cas9. Mol. Ther. Nucleic Acids 3:e216
118. Zhen S, Hua L, Takahashi Y, Narita S, Liu YH, Li Y. 2014. In vitro and in vivo growth suppression of
human papillomavirus 16-positive cervical cancer cells by CRISPR/Cas9. Biochem. Biophys. Res. Commun.
450:1422–16
119. Wang J, Quake SR. 2014. RNA-guided endonuclease provides a therapeutic strategy to cure latent
herpesviridae infection. PNAS 111:13157–62
120. Ye L, Wang J, Beyer AI, Teque F, Cradick TJ, et al. 2014. Seamless modification of wild-type induced
pluripotent stem cells to the natural CCR532 mutation confers resistance to HIV infection. PNAS
111:9591–96
121. Li C, Guan X, Du T, Jin W, Wu B, et al. 2015. Inhibition of HIV-1 infection of primary CD4+ T cells
by gene editing of CCR5 using adenovirus-delivered CRISPR/Cas9. J. Gen. Virol. 96:2381–93
122. Tebas P, Stein D, Tang WW, Frank I, Wang SQ, et al. 2014. Gene editing of CCR5 in autologous CD4
T cells of persons infected with HIV. N. Engl. J. Med. 370:901–10
123. Yi G, Choi JG, Bharaj P, Abraham S, Dang Y, et al. 2014. CCR5 gene editing of resting CD4+ T cells
by transient ZFN expression from HIV envelope pseudotyped nonintegrating lentivirus confers HIV-1
resistance in humanized mice. Mol. Ther. Nucleic Acids 3:e198
124. Badia R, Riveira-Munoz E, Clotet B, Este JA, Ballana E. 2014. Gene editing using a zinc-finger nuclease
mimicking the CCR532 mutation induces resistance to CCR5-using HIV-1. J. Antimicrob. Chemother.
69:1755–59
125. Yin H, Xue W, Chen S, Bogorad RL, Benedetti E, et al. 2014. Genome editing with Cas9 in adult mice
corrects a disease mutation and phenotype. Nat. Biotechnol. 32:551–53
126. Long C, McAnally JR, Shelton JM, Mireault AA, Bassel-Duby R, Olson EN. 2014. Prevention of mus-
cular dystrophy in mice by CRISPR/Cas9-mediated editing of germline DNA. Science 345:1184–88
127. Wu Y, Zhou H, Fan X, Zhang Y, Zhang M, et al. 2015. Correction of a genetic disease by CRISPR-
Cas9-mediated gene editing in mouse spermatogonial stem cells. Cell Res. 25:67–79
128. Wu Y, Liang D, Wang Y, Bai M, Tang W, et al. 2013. Correction of a genetic disease in mouse via use
of CRISPR-Cas9. Cell Stem Cell 13:659–62

258 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

129. Schwank G, Koo BK, Sasselli V, Dekkers JF, Heo I, et al. 2013. Functional repair of CFTR by
CRISPR/Cas9 in intestinal stem cell organoids of cystic fibrosis patients. Cell Stem Cell 13:653–58
130. Li HL, Gee P, Ishida K, Hotta A. 2015. Efficient genomic correction methods in human iPS cells using
CRISPR-Cas9 system. Methods pii:S1046–2023
131. Park CY, Kim DH, Son JS, Sung JJ, Lee J, et al. 2015. Functional correction of large factor VIII gene
chromosomal inversions in hemophilia A patient-derived iPSCs using CRISPR-Cas9. Cell Stem Cell
17:213–20
132. Xie F, Ye L, Chang JC, Beyer AI, Wang J, et al. 2014. Seamless gene correction of β-thalassemia
mutations in patient-specific iPSCs using CRISPR/Cas9 and piggyBac. Genome Res. 24:1526–33
133. Li HL, Fujimoto N, Sasakawa N, Shirai S, Ohkame T, et al. 2015. Precise correction of the dystrophin
gene in Duchenne muscular dystrophy patient induced pluripotent stem cells by TALEN and CRISPR-
Cas9. Stem Cell Rep. 4:143–54
134. Song B, Fan Y, He W, Zhu D, Niu X, et al. 2015. Improved hematopoietic differentiation efficiency of
gene-corrected β-thalassemia induced pluripotent stem cells by CRISPR/Cas9 system. Stem Cells Dev.
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

24:1053–65
135. Xu P, Tong Y, Liu XZ, Wang TT, Cheng L, et al. 2015. Both TALENs and CRISPR/Cas9 directly
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

target the HBB IVS2-654 (C > T) mutation in β-thalassemia-derived iPSCs. Sci. Rep. 5:12065
136. Smith C, Abalde-Atristain L, He C, Brodsky BR, Braunstein EM, et al. 2015. Efficient and allele-specific
genome editing of disease loci in human iPSCs. Mol. Ther. 23:570–77
137. LaFountaine JS, Fathe K, Smyth HD. 2015. Delivery and therapeutic applications of gene editing
technologies ZFNs, TALENs, and CRISPR/Cas9. Int. J. Pharm. 494:180–94
138. Wang T, Wei JJ, Sabatini DM, Lander ES. 2014. Genetic screens in human cells using the CRISPR-Cas9
system. Science 343:80–84
139. Shalem O, Sanjana NE, Hartenian E, Shi X, Scott DA, et al. 2014. Genome-scale CRISPR-Cas9 knock-
out screening in human cells. Science 343:84–87
140. Zhou Y, Zhu S, Cai C, Yuan P, Li C, et al. 2014. High-throughput screening of a CRISPR/Cas9 library
for functional genomics in human cells. Nature 509:487–91
141. Sanjana NE, Shalem O, Zhang F. 2014. Improved vectors and genome-wide libraries for CRISPR
screening. Nat. Methods 11:783–84
142. Doench JG, Hartenian E, Graham DB, Tothova Z, Hegde M, et al. 2014. Rational design of highly
active sgRNAs for CRISPR-Cas9-mediated gene inactivation. Nat. Biotechnol. 32:1262–67
143. Moore JD. 2015. The impact of CRISPR-Cas9 on target identification and validation. Drug Discov. Today
20:450–57
144. Parnas O, Jovanovic M, Eisenhaure TM, Herbst RH, Dixit A, et al. 2015. A genome-wide CRISPR
screen in primary immune cells to dissect regulatory networks. Cell 162:675–86
145. Koike-Yusa H, Li Y, Tan EP, Velasco-Herrera Mdel C, Yusa K. 2014. Genome-wide recessive genetic
screening in mammalian cells with a lentiviral CRISPR-guide RNA library. Nat. Biotechnol. 32:267–73
146. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. 1998. Potent and specific genetic
interference by double-stranded RNA in Caenorhabditis elegans. Nature 391:806–11
147. Taylor J, Woodcock S. 2015. A perspective on the future of high-throughput RNAi screening: Will
CRISPR cut out the competition or can RNAi help guide the way? J. Biomol. Screen. 20:1040–51
148. Dykxhoorn DM, Lieberman J. 2005. The silent revolution: RNA interference as basic biology, research
tool, and therapeutic. Annu. Rev. Med. 56:401–23
149. Peng R, Lin G, Li J. 2016. Potential pitfalls of CRISPR/Cas9-mediated genome editing. FEBS J.
283:1218–31
150. Fu Y, Foden JA, Khayter C, Maeder ML, Reyon D, et al. 2013. High-frequency off-target mutagenesis
induced by CRISPR-Cas nucleases in human cells. Nat. Biotechnol. 31:822–26
151. O’Geen H, Yu AS, Segal DJ. 2015. How specific is CRISPR/Cas9 really? Curr. Opin. Chem. Biol. 29:72–78
152. Wiles MV, Qin W, Cheng AW, Wang H. 2015. CRISPR-Cas9-mediated genome editing and guide
RNA design. Mamm. Genome 26:501–10
153. Guilinger JP, Thompson DB, Liu DR. 2014. Fusion of catalytically inactive Cas9 to FokI nuclease
improves the specificity of genome modification. Nat. Biotechnol. 32:577–82

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 259


BI85CH10-Qi ARI 9 May 2016 9:29

154. Tsai SQ, Wyvekens N, Khayter C, Foden JA, Thapar V, et al. 2014. Dimeric CRISPR RNA-guided
FokI nucleases for highly specific genome editing. Nat. Biotechnol. 32:569–76
155. Slaymaker IM, Gao L, Zetsche B, Scott DA, Yan WX, Zhang F. 2016. Rationally engineered Cas9
nucleases with improved specificity. Science 351:84–88
156. Fu Y, Sander JD, Reyon D, Cascio VM, Joung JK. 2014. Improving CRISPR-Cas nuclease specificity
using truncated guide RNAs. Nat. Biotechnol. 32:279–84
157. Kim D, Bae S, Park J, Kim E, Kim S, et al. 2015. Digenome-seq: genome-wide profiling of CRISPR-Cas9
off-target effects in human cells. Nat. Methods 12:237–43
158. Davis KM, Pattanayak V, Thompson DB, Zuris JA, Liu DR. 2015. Small molecule-triggered Cas9
protein with improved genome-editing specificity. Nat. Chem. Biol. 11:316–18
159. Chu VT, Weber T, Wefers B, Wurst W, Sander S, et al. 2015. Increasing the efficiency of homology-
directed repair for CRISPR-Cas9-induced precise gene editing in mammalian cells. Nat. Biotechnol.
33:543–48
160. Maruyama T, Dougan SK, Truttmann MC, Bilate AM, Ingram JR, Ploegh HL. 2015. Increasing the
efficiency of precise genome editing with CRISPR-Cas9 by inhibition of nonhomologous end joining.
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Nat. Biotechnol. 33:538–42


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

161. Lin S, Staahl B, Alla RK, Doudna JA. 2014. Enhanced homology-directed human genome engineering
by controlled timing of CRISPR/Cas9 delivery. eLife 3:e04766
162. Shrivastav M, De Haro LP, Nickoloff JA. 2008. Regulation of DNA double-strand break repair pathway
choice. Cell Res. 18:134–47
163. Yu C, Liu Y, Ma T, Liu K, Xu S, et al. 2015. Small molecules enhance CRISPR genome editing in
pluripotent stem cells. Cell Stem Cell 16:142–47
164. Liang P, Xu Y, Zhang X, Ding C, Huang R, et al. 2015. CRISPR/Cas9-mediated gene editing in human
tripronuclear zygotes. Protein Cell 6:363–72
165. Baltimore D, Berg P, Botchan M, Carroll D, Charo RA, et al. 2015. A prudent path forward for genomic
engineering and germline gene modification. Science 348:36–38
166. Lanphier E, Urnov F, Haecker SE, Werner M, Smolenski J. 2015. Don’t edit the human germ line.
Nature 519:410–11
167. Vogel G. 2015. Embryo engineering alarm. Science 347:1301
168. Kaiser J, Normile D. 2015. Embryo engineering study splits scientific community. Science 348:486–87
169. Pollack R. 2015. Eugenics lurk in the shadow of CRISPR. Science 348:871
170. Bosley KS, Botchan M, Bredenoord AL, Carroll D, Charo RA, et al. 2015. CRISPR germline
engineering—the community speaks. Nat. Biotechnol. 33:478–86
171. Mathews DJ, Chan S, Donovan PJ, Douglas T, Gyngell C, et al. 2015. CRISPR: a path through the
thicket. Nature 527:159–61
172. Esvelt KM, Smidler AL, Catteruccia F, Church GM. 2014. Concerning RNA-guided gene drives for
the alteration of wild populations. eLife 3:e03401
173. Sinkins SP, Gould F. 2006. Gene drive systems for insect disease vectors. Nat. Rev. Genet. 7:427–35
174. Larson MH, Gilbert LA, Wang X, Lim WA, Weissman JS, Qi LS. 2013. CRISPR interference
(CRISPRi) for sequence-specific control of gene expression. Nat. Protoc. 8:2180–96
175. Perez-Pinera P, Kocak DD, Vockley CM, Adler AF, Kabadi AM, et al. 2013. RNA-guided gene activation
by CRISPR-Cas9-based transcription factors. Nat. Methods 10:973–76
176. Bikard D, Jiang W, Samai P, Hochschild A, Zhang F, Marraffini LA. 2013. Programmable repression
and activation of bacterial gene expression using an engineered CRISPR-Cas system. Nucleic Acids Res.
41:7429–37
177. Konermann S, Brigham MD, Trevino AE, Hsu PD, Heidenreich M, et al. 2013. Optical control of
mammalian endogenous transcription and epigenetic states. Nature 500:472–76
178. Peters JM, Silvis MR, Zhao D, Hawkins JS, Gross CA, Qi LS. 2015. Bacterial CRISPR: accomplishments
and prospects. Curr. Opin. Microbiol. 27:121–26
179. Zalatan JG, Lee ME, Almeida R, Gilbert LA, Whitehead EH, et al. 2015. Engineering complex synthetic
transcriptional programs with CRISPR RNA scaffolds. Cell 160:339–50
180. Konermann S, Brigham MD, Trevino AE, Joung J, Abudayyeh OO, et al. 2015. Genome-scale tran-
scriptional activation by an engineered CRISPR-Cas9 complex. Nature 517:583–88

260 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

181. Shechner DM, Hacisuleyman E, Younger ST, Rinn JL. 2015. Multiplexable, locus-specific targeting of
long RNAs with CRISPR-Display. Nat. Methods 12:664–70
182. Maeder ML, Linder SJ, Cascio VM, Fu Y, Ho QH, Joung JK. 2013. CRISPR RNA-guided activation
of endogenous human genes. Nat. Methods 10:977–79
183. Cheng AW, Wang H, Yang H, Shi L, Katz Y, et al. 2013. Multiplexed activation of endogenous genes
by CRISPR-on, an RNA-guided transcriptional activator system. Cell Res. 23:1163–71
184. Chavez A, Scheiman J, Vora S, Pruitt BW, Tuttle M, et al. 2015. Highly efficient Cas9-mediated tran-
scriptional programming. Nat. Methods 12:326–28
185. Tanenbaum ME, Gilbert LA, Qi LS, Weissman JS, Vale RD. 2014. A protein-tagging system for signal
amplification in gene expression and fluorescence imaging. Cell 159:635–46
186. He S, Weintraub SJ. 1998. Stepwise recruitment of components of the preinitiation complex by upstream
activators in vivo. Mol. Cell. Biol. 18:2876–83
187. Govind CK, Yoon S, Qiu H, Govind S, Hinnebusch AG. 2005. Simultaneous recruitment of coactivators
by Gcn4p stimulates multiple steps of transcription in vivo. Mol. Cell. Biol. 25:5626–38
188. Hawkins JS, Wong S, Peters JM, Almeida R, Qi LS. 2015. Targeted transcriptional repression in bacteria
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

using CRISPR interference (CRISPRi). Methods Mol. Biol. 1311:349–62


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

189. Polstein L, Perez-Pinera P, Kocak D, Vockley C, Bledsoe P, et al. 2015. Genome-wide specificity of
DNA-binding, gene regulation, and chromatin remodeling by TALE- and CRISPR/Cas9-based tran-
scriptional activators. Genome Res. 25:1158–69
190. Zhao Y, Dai Z, Liang Y, Yin M, Ma K, et al. 2014. Sequence-specific inhibition of microRNA via
CRISPR/CRISPRi system. Sci. Rep. 4:3943
191. La Russa MF, Qi LS. 2015. The new state of the art: Cas9 for gene activation and repression. Mol. Cell.
Biol. 35:3800–9
192. Dominguez AA, Lim WA, Qi LS. 2016. Beyond editing: repurposing CRISPR-Cas9 for precision
genome regulation and interrogation. Nat. Rev. Mol. Cell Biol. 17:5–15
193. Lin S, Ewen-Campen B, Ni X, Housden BE, Perrimon N. 2015. In vivo transcriptional activation using
CRISPR/Cas9 in Drosophila. Genetics 201:433–42
194. Piatek A, Ali Z, Baazim H, Li L, Abulfaraj A, et al. 2015. RNA-guided transcriptional regulation in planta
via synthetic dCas9-based transcription factors. Plant Biotechnol. J. 13:578–89
195. Lowder LG, Zhang D, Baltes NJ, Paul JW 3rd, Tang X, et al. 2015. A CRISPR/Cas9 toolbox for
multiplexed plant genome editing and transcriptional regulation. Plant Physiol. 169:971–85
196. Zhang Y, Yin C, Zhang T, Li F, Yang W, et al. 2015. CRISPR/gRNA-directed synergistic activation
mediator (SAM) induces specific, persistent and robust reactivation of the HIV-1 latent reservoirs. Sci.
Rep. 5:16277
197. Mendenhall EM, Williamson KE, Reyon D, Zou JY, Ram O, et al. 2013. Locus-specific editing of
histone modifications at endogenous enhancers. Nat. Biotechnol. 31:1133–36
198. Snowden AW, Gregory PD, Case CC, Pabo CO. 2002. Gene-specific targeting of H3K9 methylation
is sufficient for initiating repression in vivo. Curr. Biol. 12:2159–66
199. Maeder ML, Angstman JF, Richardson ME, Linder SJ, Cascio VM, et al. 2013. Targeted DNA
demethylation and activation of endogenous genes using programmable TALE-TET1 fusion proteins.
Nat. Biotechnol. 31:1137–42
200. Rivenbark AG, Stolzenburg S, Beltran AS, Yuan X, Rots MG, et al. 2012. Epigenetic reprogramming
of cancer cells via targeted DNA methylation. Epigenetics 7:350–60
201. Konermann S, Brigham MD, Trevino AE, Hsu PD, Heidenreich M, et al. 2013. Optical control of
mammalian endogenous transcription and epigenetic states. Nature 500:472–76
202. Keung AJ, Bashor CJ, Kiriakov S, Collins JJ, Khalil AS. 2014. Using targeted chromatin regulators to
engineer combinatorial and spatial transcriptional regulation. Cell 158:110–20
203. Thakore PI, D’Ippolito AM, Song L, Safi A, Shivakumar NK, et al. 2015. Highly specific epigenome
editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements. Nat. Methods 12:1143–49
204. Lanctôt C, Cheutin T, Cremer M, Cavalli G, Cremer T. 2007. Dynamic genome architecture in the
nuclear space: regulation of gene expression in three dimensions. Nat. Rev. Genet. 8:104–15
205. Schneider R, Grosschedl R. 2007. Dynamics and interplay of nuclear architecture, genome organization,
and gene expression. Genes Dev. 21:3027–43

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 261


BI85CH10-Qi ARI 9 May 2016 9:29

206. Dixon JR, Jung I, Selvaraj S, Shen Y, Antosiewicz-Bourget JE, et al. 2015. Chromatin architecture
reorganization during stem cell differentiation. Nature 518:331–36
207. Peric-Hupkes D, Meuleman W, Pagie L, Bruggeman SW, Solovei I, et al. 2010. Molecular maps of the
reorganization of genome–nuclear lamina interactions during differentiation. Mol. Cell 38:603–13
208. Gall JG, Pardue ML. 1969. Formation and detection of RNA–DNA hybrid molecules in cytological
preparations. PNAS 63:378–83
209. John HA, Birnstiel ML, Jones KW. 1969. RNA–DNA hybrids at the cytological level. Nature 223:582–87
210. Pardue ML, Gall JG. 1969. Molecular hybridization of radioactive DNA to the DNA of cytological
preparations. PNAS 64:600–4
211. Pinkel D, Straume T, Gray JW. 1986. Cytogenetic analysis using quantitative, high-sensitivity, fluores-
cence hybridization. PNAS 83:2934–38
212. Pinkel D, Gray JW, Trask B, van den Engh G, Fuscoe J, van Dekken H. 1986. Cytogenetic analysis by
in situ hybridization with fluorescently labeled nucleic acid probes. Cold Spring Harb. Symp. Quant. Biol.
51(Pt 1):151–57
213. Speicher MR, Gwyn Ballard S, Ward DC. 1996. Karyotyping human chromosomes by combinatorial
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

multi-fluor FISH. Nat. Genet. 12:368–75


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

214. Vorsanova SG, Yurov YB, Iourov IY. 2010. Human interphase chromosomes: a review of available
molecular cytogenetic technologies. Mol. Cytogenet. 3:1
215. Riegel M. 2014. Human molecular cytogenetics: from cells to nucleotides. Genet. Mol. Biol. 37:194–209
216. Cmarko D, Ligasova A, Koberna K. 2014. Tracking DNA and RNA sequences at high resolution. Methods
Mol. Biol. 1117:343–66
217. Hutchison NJ, Langer-Safer PR, Ward DC, Hamkalo BA. 1982. In situ hybridization at the electron
microscope level: hybrid detection by autoradiography and colloidal gold. J. Cell Biol. 95:609–18
218. Molenaar C, Wiesmeijer K, Verwoerd NP, Khazen S, Eils R, et al. 2003. Visualizing telomere dynamics
in living mammalian cells using PNA probes. EMBO J. 22:6631–41
219. Robinett CC, Straight A, Li G, Willhelm C, Sudlow G, et al. 1996. In vivo localization of DNA sequences
and visualization of large-scale chromatin organization using lac operator/repressor recognition. J. Cell
Biol. 135:1685–700
220. Marshall WF, Straight A, Marko JF, Swedlow J, Dernburg A, et al. 1997. Interphase chromosomes
undergo constrained diffusional motion in living cells. Curr. Biol. 7:930–39
221. Li P, Jin H, Hoang ML, Yu HG. 2011. Tracking chromosome dynamics in live yeast cells: coordinated
movement of rDNA homologs and anaphase disassembly of the nucleolus during meiosis. Chromosome
Res. 19:1013–26
222. Wang X, Kam Z, Carlton PM, Xu L, Sedat JW, Blackburn EH. 2008. Rapid telomere motions in live
human cells analyzed by highly time-resolved microscopy. Epigenetics Chromatin 1:4
223. Shelby RD, Hahn KM, Sullivan KF. 1996. Dynamic elastic behavior of α-satellite DNA domains visu-
alized in situ in living human cells. J. Cell Biol. 135:545–57
224. Bronstein I, Israel Y, Kepten E, Mai S, Shav-Tal Y, et al. 2009. Transient anomalous diffusion of
telomeres in the nucleus of mammalian cells. Phys. Rev. Lett. 103:018102
225. Anton T, Bultmann S, Leonhardt H, Markaki Y. 2014. Visualization of specific DNA sequences in living
mouse embryonic stem cells with a programmable fluorescent CRISPR/Cas system. Nucleus 5:163–72
226. Deng W, Shi X, Tjian R, Lionnet T, Singer RH. 2015. CASFISH: CRISPR/Cas9-mediated in situ
labeling of genomic loci in fixed cells. PNAS 112:11870–75
227. Kuscu C, Arslan S, Singh R, Thorpe J, Adli M. 2014. Genome-wide analysis reveals characteristics of
off-target sites bound by the Cas9 endonuclease. Nat. Biotechnol. 32:677–83
228. Duan J, Lu G, Xie Z, Lou M, Luo J, et al. 2014. Genome-wide identification of CRISPR/Cas9 off-targets
in human genome. Cell Res. 24:1009–12
229. Dow LE, Fisher J, O’Rourke KP, Muley A, Kastenhuber ER, et al. 2015. Inducible in vivo genome
editing with CRISPR-Cas9. Nat. Biotechnol. 33:390–94
230. Gonzalez F, Zhu Z, Shi ZD, Lelli K, Verma N, et al. 2014. An iCRISPR platform for rapid, multiplexable,
and inducible genome editing in human pluripotent stem cells. Cell Stem Cell 15:215–26
231. Nihongaki Y, Yamamoto S, Kawano F, Suzuki H, Sato M. 2015. CRISPR-Cas9-based photoactivatable
transcription system. Chem. Biol. 22:169–74

262 Wang · La Russa · Qi


BI85CH10-Qi ARI 9 May 2016 9:29

232. Polstein LR, Gersbach CA. 2015. A light-inducible CRISPR-Cas9 system for control of endogenous
gene activation. Nat. Chem. Biol. 11:198–200
233. Zetsche B, Volz SE, Zhang F. 2015. A split-Cas9 architecture for inducible genome editing and tran-
scription modulation. Nat. Biotechnol. 33:139–42
234. Wright AV, Sternberg SH, Taylor DW, Staahl BT, Bardales JA, et al. 2015. Rational design of a split-
Cas9 enzyme complex. PNAS 112:2984–89
235. Nihongaki Y, Kawano F, Nakajima T, Sato M. 2015. Photoactivatable CRISPR-Cas9 for optogenetic
genome editing. Nat. Biotechnol. 33:755–60
236. Truong DJ, Kuhner K, Kuhn R, Werfel S, Engelhardt S, et al. 2015. Development of an intein-mediated
split-Cas9 system for gene therapy. Nucleic Acids Res. 43:6450–58
237. Sampson TR, Weiss DS. 2013. Cas9-dependent endogenous gene regulation is required for bacterial
virulence. Biochem. Soc. Trans. 41:1407–11
238. Sampson TR, Saroj SD, Llewellyn AC, Tzeng YL, Weiss DS. 2013. A CRISPR/Cas system mediates
bacterial innate immune evasion and virulence. Nature 497:254–57
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

239. Nelles DA, Fang MY, Aigner S, Yeo GW. 2015. Applications of Cas9 as an RNA-programmed RNA-
binding protein. BioEssays 37:732–39
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

240. Ma H, Wu Y, Dang Y, Choi JG, Zhang J, Wu H. 2014. Pol III promoters to express small RNAs:
delineation of transcription initiation. Mol. Ther. Nucleic Acids 3:e161
241. Hwang WY, Fu Y, Reyon D, Maeder ML, Kaini P, et al. 2013. Heritable and precise zebrafish genome
editing using a CRISPR-Cas system. PLOS ONE 8:e68708
242. Nielsen S, Yuzenkova Y, Zenkin N. 2013. Mechanism of eukaryotic RNA polymerase III transcription
termination. Science 340:1577–80
243. Pengue G, Lania L. 1996. Krüppel-associated box-mediated repression of RNA polymerase II promoters
is influenced by the arrangement of basal promoter elements. PNAS 93:1015–20
244. Groner AC, Meylan S, Ciuffi A, Zangger N, Ambrosini G, et al. 2010. KRAB–zinc finger proteins and
KAP1 can mediate long-range transcriptional repression through heterochromatin spreading. PLOS
Genet. 6:e1000869
245. Mittler G, Stühler T, Santolin L, Uhlmann T, Kremmer E, et al. 2003. A novel docking site on Mediator
is critical for activation by VP16 in mammalian cells. EMBO J. 22:6494–504
246. Gaj T, Epstein BE, Schaffer DV. 2016. Genome engineering using adeno-associated virus: basic and
clinical research applications. Mol. Ther. 24:458–64
247. Friedland AE, Baral R, Singhal P, Loveluck K, Shen S, et al. 2015. Characterization of Staphylococcus
aureus Cas9: a smaller Cas9 for all-in-one adeno-associated virus delivery and paired nickase applications.
Genome Biol. 16:257
248. Crispo M, Mulet AP, Tesson L, Barrera N, Cuadro F, et al. 2015. Efficient generation of myostatin knock-
out sheep using CRISPR/Cas9 technology and microinjection into zygotes. PLOS ONE 10:e0136690
249. Bassett AR, Tibbit C, Ponting CP, Liu JL. 2013. Highly efficient targeted mutagenesis of Drosophila
with the CRISPR/Cas9 system. Cell Rep. 4:220–28
250. Xue W, Chen S, Yin H, Tammela T, Papagiannakopoulos T, et al. 2014. CRISPR-mediated direct
mutation of cancer genes in the mouse liver. Nature 514:380–84
251. Cho SW, Lee J, Carroll D, Kim JS. 2013. Heritable gene knockout in Caenorhabditis elegans by direct
injection of Cas9-sgRNA ribonucleoproteins. Genetics 195:1177–80
252. Sung YH, Kim JM, Kim HT, Lee J, Jeon J, et al. 2014. Highly efficient gene knockout in mice and
zebrafish with RNA-guided endonucleases. Genome Res. 24:125–31
253. Liu J, Gaj T, Yang Y, Wang N, Shui S, et al. 2015. Efficient delivery of nuclease proteins for genome
editing in human stem cells and primary cells. Nat. Protoc. 10:1842–59
254. Schumann K, Lin S, Boyer E, Simeonov DR, Subramaniam M, et al. 2015. Generation of knock-in
primary human T cells using Cas9 ribonucleoproteins. PNAS 112:10437–42
255. Hendel A, Bak RO, Clark JT, Kennedy AB, Ryan DE, et al. 2015. Chemically modified guide RNAs
enhance CRISPR-Cas genome editing in human primary cells. Nat. Biotechnol. 33:985–89
256. Liang X, Potter J, Kumar S, Zou Y, Quintanilla R, et al. 2015. Rapid and highly efficient mammalian
cell engineering via Cas9 protein transfection. J. Biotechnol. 208:44–53

www.annualreviews.org • CRISPR/Cas9 in Genome Editing and Beyond 263


BI85CH10-Qi ARI 9 May 2016 9:29

257. Zuris JA, Thompson DB, Shu Y, Guilinger JP, Bessen JL, et al. 2015. Cationic lipid-mediated delivery
of proteins enables efficient protein-based genome editing in vitro and in vivo. Nat. Biotechnol. 33:73–80
258. Ramakrishna S, Kwaku Dad AB, Beloor J, Gopalappa R, Lee SK, Kim H. 2014. Gene disruption by
cell-penetrating peptide-mediated delivery of Cas9 protein and guide RNA. Genome Res. 24:1020–27
259. D’Astolfo DS, Pagliero RJ, Pras A, Karthaus WR, Clevers H, et al. 2015. Efficient intracellular delivery
of native proteins. Cell 161:674–90
260. Sun W, Ji W, Hall JM, Hu Q, Wang C, et al. 2015. Self-assembled DNA nanoclews for the efficient
delivery of CRISPR-Cas9 for genome editing. Angew. Chem. Int. Ed. Engl. 54:12029–33
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

264 Wang · La Russa · Qi


BI85-FrontMatter ARI 20 May 2016 8:42

Annual Review of
Biochemistry
Contents Volume 85, 2016

Cellular Homeostasis and Aging


F. Ulrich Hartl p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Dietary Protein, Metabolism, and Aging
George A. Soultoukis and Linda Partridge p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 5
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

Signaling Networks Determining Life Span


Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Celine E. Riera, Carsten Merkwirth, C. Daniel De Magalhaes Filho,


and Andrew Dillin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p35
Mitochondrial Gene Expression: A Playground of Evolutionary
Tinkering
Walter Neupert p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p65
Organization and Regulation of Mitochondrial Protein Synthesis
Martin Ott, Alexey Amunts, and Alan Brown p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p77
Structure and Function of the Mitochondrial Ribosome
Basil J. Greber and Nenad Ban p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 103
Maintenance and Expression of Mammalian Mitochondrial DNA
Claes M. Gustafsson, Maria Falkenberg, and Nils-Göran Larsson p p p p p p p p p p p p p p p p p p p p p p 133
Enjoy the Trip: Calcium in Mitochondria Back and Forth
Diego De Stefani, Rosario Rizzuto, and Tullio Pozzan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 161
Mechanics and Single-Molecule Interrogation of DNA Recombination
Jason C. Bell and Stephen C. Kowalczykowski p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 193
CRISPR/Cas9 in Genome Editing and Beyond
Haifeng Wang, Marie La Russa, and Lei S. Qi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 227
Nucleotide Excision Repair and Transcriptional Regulation: TFIIH
and Beyond
Emmanuel Compe and Jean-Marc Egly p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 265
Transcription as a Threat to Genome Integrity
Hélène Gaillard and Andrés Aguilera p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 291
Mechanisms of Bacterial Transcription Termination: All Good Things
Must End
Ananya Ray-Soni, Michael J. Bellecourt, and Robert Landick p p p p p p p p p p p p p p p p p p p p p p p p p p p p 319

v
BI85-FrontMatter ARI 20 May 2016 8:42

Nucleic Acid–Based Nanodevices in Biological Imaging


Kasturi Chakraborty, Aneesh T. Veetil, Samie R. Jaffrey, and Yamuna Krishnan p p p p p 349
The p53 Pathway: Origins, Inactivation in Cancer, and Emerging
Therapeutic Approaches
Andreas C. Joerger and Alan R. Fersht p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 375
The Substrate Specificity of Sirtuins
Poonam Bheda, Hui Jing, Cynthia Wolberger, and Hening Lin p p p p p p p p p p p p p p p p p p p p p p p p p 405
Macrodomains: Structure, Function, Evolution, and Catalytic Activities
Johannes Gregor Matthias Rack, Dragutin Perina, and Ivan Ahel p p p p p p p p p p p p p p p p p p p p p p 431
Biosynthesis of the Metalloclusters of Nitrogenases
Yilin Hu and Markus W. Ribbe p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 455
Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Radical S-Adenosylmethionine Enzymes in Human Health and Disease


Bradley J. Landgraf, Erin L. McCarthy, and Squire J. Booker p p p p p p p p p p p p p p p p p p p p p p p p p p 485
Ice-Binding Proteins and Their Function
Maya Bar Dolev, Ido Braslavsky, and Peter L. Davies p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 515
Shared Molecular Mechanisms of Membrane Transporters
David Drew and Olga Boudker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 543
Spatial and Temporal Regulation of Receptor Tyrosine Kinase
Activation and Intracellular Signal Transduction
John J.M. Bergeron, Gianni M. Di Guglielmo, Sophie Dahan,
Michel Dominguez, and Barry I. Posner p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 573
Understanding the Chemistry and Biology of Glycosylation with
Glycan Synthesis
Larissa Krasnova and Chi-Huey Wong p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 599
The Biochemistry of O-GlcNAc Transferase: Which Functions Make
It Essential in Mammalian Cells?
Zebulon G. Levine and Suzanne Walker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 631
Mechanisms of Mitotic Spindle Assembly
Sabine Petry p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 659
Mammalian Autophagy: How Does It Work?
Carla F. Bento, Maurizio Renna, Ghita Ghislat, Claudia Puri, Avraham Ashkenazi,
Mariella Vicinanza, Fiona M. Menzies, and David C. Rubinsztein p p p p p p p p p p p p p p p p p p 685
Experimental Milestones in the Discovery of Molecular Chaperones
as Polypeptide Unfolding Enzymes
Andrija Finka, Rayees U.H. Mattoo, and Pierre Goloubinoff p p p p p p p p p p p p p p p p p p p p p p p p p p p p 715
Necroptosis and Inflammation
Kim Newton and Gerard Manning p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 743

vi Contents
BI85-FrontMatter ARI 20 May 2016 8:42

Reactive Oxygen Species and Neutrophil Function


Christine C. Winterbourn, Anthony J. Kettle, and Mark B. Hampton p p p p p p p p p p p p p p p p p p 765

Indexes

Cumulative Index of Contributing Authors, Volumes 81–85 p p p p p p p p p p p p p p p p p p p p p p p p p p p 793


Cumulative Index of Article Titles, Volumes 81–85 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 797

Errata

An online log of corrections to Annual Review of Biochemistry articles may be found at


Annu. Rev. Biochem. 2016.85:227-264. Downloaded from www.annualreviews.org

http://www.annualreviews.org/errata/biochem
Access provided by 186.46.223.230 on 06/03/20. For personal use only.

Contents vii

You might also like